tribological studies on the surface of glasses by lateral

127

Upload: others

Post on 25-Dec-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Tribological studies on the surface of glasses by lateral

Tribological studies on the surface of glassesby lateral indentation technique

Dissertation

zur Erlangung des akademischen Grades Doktor-Ingenieur

(Dr.-Ing.)

vorgelegt dem Rat der Chemisch-Geowissenschaftlichen Fakultaet

der Friedrich-Schiller-Universitaet Jena

Elham Moayedi

Geboren am 17.01.1982 in Teheran

Page 2: Tribological studies on the surface of glasses by lateral

Dissertation, Friedrich-Schiller-Universität Jena, [2021]

Reviewers:

1. Prof. Dr. Ing. Lothar Wondraczek, Otto Schott Institute of Materials

Research, Friedrich Schiller University of Jena

2. Prof. Morten Mattrup Smedskjaer, Department of Chemistry and Bio-

science, Aalborg University

Date of defense: 14.07.2021

Page 3: Tribological studies on the surface of glasses by lateral

Acknowledgements

First and foremost, I would like to express my sincere gratitude to Prof.

Dr. Lothar Wondraczek for his guidance throughout this thesis, for valuable

discussions and for providing outstanding opportunities for collaborations.

By the chance he gave me, I noticed how much I love research and learning

new skills and topics.

I gratefully acknowledge �nancial support from the European Research

Council (ERC) under the European Union's Horizon 2020 research and inno-

vation program (ERC grant UTOPES, grant agreement no. 681652) as well

as Starting Ramp scheme of Priority Program of the German Science Foun-

dation, �Topological Engineering of Ultra-Strong Glasses�(DFG SPP1594).

Within said ERC grant UTOPES; my work pro�ted immensely from collab-

oration with Prof. Dr. Enrico Gnecco. I am very grateful for his support

and discussions. Special thanks to Jana Henning. As part of her own Ph.D.

thesis she carried out AFM measurements.

Furthermore, I would like to thank my colleague René Limbach for valu-

able help with indentation experiments, inspiring discussions and outline of

this thesis, Peter for the hours of calculations and estimations of some data,

Ferdinand (former) for the support with OriginLab, Jan with image process-

ing, Gohar with literature providing, Shigeki (former) for all the precious

scienti�c discussions and reviewing my publication. For all scienti�c discus-

sions and the good time that I spent in the institute, I would like to thank my

current and former colleagues Bruno, Guilherme, Benjamin, Aaron, Kristin,

Ali, Lenka, Theresia, Huyen, Courtney, Garth, Caio, Doris, Vivi, Yuko, Fe-

lix, Byoungjin, Lingqi, Yang, Xu, Ru, Atef, Ding, Pigter, Roman, Omar,

Jelena, Vahid, Ayda, Aziz, Thien, and Michal. My sincere gratitude to the

kindest secretary (former) Ute Böttger, and technicians Christian Zeidler,

3

Page 4: Tribological studies on the surface of glasses by lateral

4 Gutachter

Gabriele Möller (former), Nadja Büchert, Thomas Kittel (former) and Clau-

dia Siedler, for all their kindness and e�ort to support my studies.

I would also like to state my appreciation to the external collaborators

at the Leibniz Institute of Photonic Technology (IPHT), Dr. Jan Delith and

Andrea Delith for performing AFM measurements for me.

Finally, my warm gratitude to all my family and friends. Specially the

ones that helped me by their kind advice and encouragements during these

di�cult times of pandemic. I am very grateful that my beloved husband

Enrique Fernandez helped me with the Latex issues and stayed with lots of

patience on my side.

Page 5: Tribological studies on the surface of glasses by lateral

ت رون ز ه ود آ ن و ا

ق ت و ن ت دا س

د ودا ی از س

ت د ی دا س ت زآن روی

یام

Amidst this Strait which swells of Chasm’s Hide,

No salt would dare his shaky vessel guide!

Upon say-so each captain mapped a chart,

But none could tell what lies beyond the tide!

Omar Khayyam

Translated by Edward FitzGerald

5

Page 6: Tribological studies on the surface of glasses by lateral
Page 7: Tribological studies on the surface of glasses by lateral

Contents

Contents viii

List of Figures xii

List of Tables xiii

Abstract 1

Glossary 3

1 Introduction 7

2 State of the art 11

2.1 Deformation in glasses . . . . . . . . . . . . . . . . . . . . . . 11

2.2 Formation of cracks in contact-induced damaged surface . . . 13

2.3 Parameters a�ecting deformation regimes . . . . . . . . . . . 15

2.3.1 The indenter geometry . . . . . . . . . . . . . . . . . . 15

2.3.2 The applied normal force . . . . . . . . . . . . . . . . 16

2.3.3 Scratching rate . . . . . . . . . . . . . . . . . . . . . . 17

2.3.4 The glass surface condition . . . . . . . . . . . . . . . 18

2.3.5 Environmental atmosphere and humidity . . . . . . . . 20

2.4 Compositional dependence of deformation in glass . . . . . . . 22

2.5 Contribution of friction to materials behaviour in indentation

experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

vii

Page 8: Tribological studies on the surface of glasses by lateral

viii CONTENTS

3 Experimental methodology 29

3.1 Ramp load scratching . . . . . . . . . . . . . . . . . . . . . . 29

3.1.1 Ramp load scratching test on vitreous silica . . . . . . 29

3.1.2 Scratching of metallic glass by a spherical indenter . . 32

3.2 Constant low load scratching . . . . . . . . . . . . . . . . . . 33

3.2.1 Scratching at constant load by a Berkovich indenter

on vitreous silica . . . . . . . . . . . . . . . . . . . . . 33

3.2.2 Constant low load scratching on silicate glasses . . . . 34

3.2.2.1 Glass samples . . . . . . . . . . . . . . . . . 34

3.2.2.2 General characterization . . . . . . . . . . . . 34

3.2.2.3 Scratching and indentation tests . . . . . . . 35

3.2.2.4 AFM imaging and subsequent heat treatment 35

4 Results and discussions 37

4.1 Studies at low load scratching in plastic regime of deformation 39

4.1.1 Relaxation of scratch-induced surface deformation in

silicate glasses . . . . . . . . . . . . . . . . . . . . . . . 39

4.1.2 Rippling inside the scratch groove of vitreous silica . . 47

4.2 Studies at ramp load scratching in plastic regime of deformation 56

4.2.1 Statistical analysis of microabrasion onset in vitreous

silica . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

4.2.2 Scratching of metallic glass by a blunt indenter . . . . 67

5 Conclusions 79

Zusammenfassung 83

Bibliography 108

Appendix 109

Selbstsändigkeitserklärung 112

Page 9: Tribological studies on the surface of glasses by lateral

List of Figures

2.1 Typical scratch pattern made on the surface of the Soda lime

silica glass by a vickers penetrator (leading edge) during a

monotonic loading cycle [1]. . . . . . . . . . . . . . . . . . . . 14

2.2 Maximum density relative variation after high hydrostatic pres-

sure testing at room temperature as a function of poisson's ratio. 23

3.1 Schematic of the determination of lateral load during lateral

displacement of Berkovich indenter . . . . . . . . . . . . . . . 30

3.2 a 3D representation of the employed Berkovich tip, obtained

by wide-�eld confocal microscopy. . . . . . . . . . . . . . . . . 31

3.3 The Berkovich tip edge-forward and face-forward con�gura-

tions during scanning . . . . . . . . . . . . . . . . . . . . . . . 31

4.1 (a) Typical scratch pattern which is observed on silicate glasses

during steady scratching with increasing normal load. (b)

Scratch pattern of vitreous silica observed at ramping load of

0.05-300 mN and scratching rate of 100 µm/s. . . . . . . . . . 38

4.2 Typical AFM micrograph of a scratch groove as used for vol-

ume analysis for boro�oat 33 . . . . . . . . . . . . . . . . . . 40

4.3 AFM top-views and cross-pro�le scans of the residual imprints

after Berkovich indentation, before and after annealing for 1

h at 0.95Tg, shown for silica, BF33, and SLG. . . . . . . . . . 41

4.4 AFM top-views and cross-pro�le scans of the residual scratch

grooves after Berkovich edge-forward scratching, before and

after annealing at 0.95Tg, shown for silica, BF33, and SLG. . 44

ix

Page 10: Tribological studies on the surface of glasses by lateral

x LIST OF FIGURES

4.5 (a) Cross-pro�le data of scratches on SiO2 before and after

annealing at 0.95Tg for 1 and 2h. (b) Recovery ratio versus

annealing time for di�erent scratch rates. . . . . . . . . . . . 45

4.6 Cross-pro�le scans over a scratch groove on BF33 before and

after 1h annealing at 0.95Tg for face-forward and edge-forward

orientation of the Berkovich tip. . . . . . . . . . . . . . . . . . 46

4.7 Recovery ratios as a function of (a) loading values and (b)

Poisson's ratio for normal and anomalous glasses. . . . . . . . 47

4.8 An atomic force microscopy image of a part of one scratch

performed at scratching rate of 10 µm and increasing normal

load of 10 mN showing ripples in the scratch groove. . . . . . 48

4.9 (a) AFM topography of a silica glass surface previously scratched

with a normal force of 30 mN and a scan velocity of 10 µm/s.

Set point: FN = 1.8 mN. (b) Cross section along the light

blue line in (a). (c) Simulated herringbone pattern obtained

from the simple repetition of Berkovich geometry. (b) Cross

section along the light blue line in (a). (c) Simulated herring-

bone pattern obtained from the simple repetition of Berkovich

geometry every 350 nm without relaxation e�ects. . . . . . . . 50

4.10 (a) AFM error signal across the wear scar in Fig. 4.9(a). (b)

2D self-correlation and (c) 1D-FFT along longitudinal direc-

tion extracted from the region corresponding to the herring-

bone pattern in (a); (d), (e) cross sections along the blue lines

in (b) and (c), respectively. . . . . . . . . . . . . . . . . . . . 51

4.11 Laser scanning microscopy image of a wear groove obtained

in the same conditions of Fig. 4.9(a). . . . . . . . . . . . . . . 52

4.12 Velocity dependence of the ripples period as measured by

AFM ex situ after scratching with a normal force FN = 30

mN (blue dots) and linear �t of the experimental data points

(red curve). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

4.13 (a) Time variation of the indentation depth while scratching

with v = 10 µm/s. (b) Lateral force vs. indenter position

when the scratching process initiates (at x = 20 µm). . . . . . 54

Page 11: Tribological studies on the surface of glasses by lateral

LIST OF FIGURES xi

4.14 Evolution of lateral force during scratching while increasing

the normal load from 0.05 mN to 300 mN at a scratching

velocity of 500 µm/s. . . . . . . . . . . . . . . . . . . . . . . . 58

4.15 (a) Scratch pattern for fused silica at a scratching rate of 50

µm/s and a normal load which increases from zero to 300

mN, using an irregular diamond edge for scratching. (b) is a

representation of the corresponding variation in the apparent

friction coe�cient. . . . . . . . . . . . . . . . . . . . . . . . . 59

4.16 Determination of the onset of microabrasion (scratch length in

µm) OM through di�erent methods: (a) Post mortem optical

microscopy and in-situ observation of the apparent coe�cient

of friction, and (b) optical microscopy and in-situ observation

of the lateral force. In (c) the determination of lateral force

is considered, i.e., as read directly during in-situ scans and

as determined from the length at which OM was observed

through the apparent coe�cient of friction, µ, according to (a). 62

4.17 Post mortem optical microscopic image of a scratch generated

at a scratching speed of 10 µm/s under increasing normal load

(3 mN/s). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

4.18 Statistical analysis of the onset of microabrasion (OM) in vit-

reous silica during lateral indentation. . . . . . . . . . . . . . 65

4.19 (a) An Scanning Electron Microscopy overview image of two

scratches performed at load of 30 mN and scratching rate of

10 µm/s. The whole scratching length is not shown in the

image. (b) A pro�le of displacement into surface over the

whole lateral length for a scratch at the same conditions. . . . 69

4.20 The normal load vs. penetratuin depth curves for indentation

(left) and scratching (right) experiments at normal load of 20,

30 and 40 mN. All scratching experiments were performed at

scratching rate of 10 µm/s. . . . . . . . . . . . . . . . . . . . 70

4.21 Normal loads at �rst pop-ins appearance for indentations and

scratching (left) and lateral loads for scratching experiments

(right). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73

Page 12: Tribological studies on the surface of glasses by lateral

xii LIST OF FIGURES

4.22 Coe�cients of frictions for scratches at loads of 20, 30, 40 and

50 mN. Each experiment was repeated 5 times. . . . . . . . . 74

4.23 a) An image of a scratch at the load of 30 mN obtained by

Scanning Electron Microscope b) and c) Atomic force mi-

croscopy images of two marked sections of the same scratch in

(a), d) and e) Center line pro�les of AFM images in (c) and

(d). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76

4.24 Comparison of pop-in loads read from indenter and the point

of appearance of patterns in SEM images for the scratching

loads of 20 and 30 mN. An standard deviation of 5% µm

should be taken into account for the data. . . . . . . . . . . . 77

F1 (a) Width and (b) depth of the wear grooves formed on a

silica glass surface scratched by a Berkovich diamond tip with

a scan velocity of 10 µm/s and di�erent normal loads. . . . . 110

F2 (a) AFM topography corresponding to Fig. 4.10(a); (b) hori-

zontal cross-section along the blue line in (a). . . . . . . . . . 110

F3 AFM topography of the very end of the scratch. Frame size:

14.1 µm∗6.1 µm. . . . . . . . . . . . . . . . . . . . . . . . . . 111

F4 (a) AFM topography of one section of scratch performed at

load of 30 mN. (b) Pro�le along the scratch shown with a

horizontal line in (a). (c) FFT analysis of the same image.

(d) Pro�le along the line in (c). . . . . . . . . . . . . . . . . . 111

Page 13: Tribological studies on the surface of glasses by lateral

List of Tables

3.1 Glass transition temperature Tg, density ρ, Young's modulus

E, shear modulus G, bulk modulus K, atomic packing density

Cg and Poission's ratio ν of the studied glasses. . . . . . . . . 34

4.1 Indentation and scratch volumes before and after annealing

for di�erent experimental conditions. . . . . . . . . . . . . . . 42

4.2 Onset of microabrasion (OM) for a series of 20 experiments,

scratching vitreous silica at a rate of 50 µm/s with normal

load increasing from 0.05 mN at a rate of 15 mN/s. . . . . . . 64

4.3 Weibull parameters for failure modes I and II and varying

speed of scratching. . . . . . . . . . . . . . . . . . . . . . . . . 66

4.4 Normal loads and lateral loads at which the pop-ins appear

for each experimental load at 20, 30, 40, and 50 mN. . . . . . 72

xiii

Page 14: Tribological studies on the surface of glasses by lateral
Page 15: Tribological studies on the surface of glasses by lateral

Abstract

Glass is a material particularly sensitive to surface damage when exposed to

abrasive loads and scratching. Such local mechanical contacts not only com-

promise the surface quality and mechanical performance, but also degrade

the visual appearance of material. Understanding abrasive damage and the

underlying material properties has therefore been a subject of signi�cant in-

terest. Hence instrumented lateral nanoindentation has been employed in

this thesis to obtain quantitative and quantitative information on the onset

of scratch-induced surface of three silicate glasses and one metallic glass in

low load and high load regimes. First, the role of compaction and shear �ow

in the deformation caused by lateral indentations on silicate glasses is quan-

ti�ed through classical relaxation experiments. In the anomalous glasses,

the main mechanism of plastic deformation in lateral indentations was re-

vealed to be densi�cation, while in the normal glasses, shear �ow played a

considerable role. Furthermore, lateral indentation was used to study the

appearance of wavy patterns formed by sliding on a compliant surface which

was investigated later by post mortem AFM technique. The average repeti-

tion distance of the ripples which is in the sub-µm range was displayed to be

dependent on the scratching velocity. Additionally, a correlation was noticed

between scratching velocity (loading rate) and the formation of microabra-

sion regime in vitreous silica when analysing the data through Weibull distri-

bution. The quantitative information obtained through Weibull distribution

analysis showed the considerable role of tested volume in the appearance of

di�erent types of �aws. Not only scratching velocity, but also loading can in-

�uence the deformation behaviour of glasses. This was further exhibited by

analysis of pop-in loads when performing lateral indentations on the surface

of a metallic glass.

1

Page 16: Tribological studies on the surface of glasses by lateral
Page 17: Tribological studies on the surface of glasses by lateral

Glossary

Cg Atomic packing density. 32

Er combined elastic response of the indenter tip and the glass specimen. 56

FN Normal load. 37

Ff Coulumbian friction e�ort. 25

Fk Kinetic friction force. 46

Fs Threshold value for friction force. 47

Ft Tangential force. 25

Fcr Cracking e�ort. 25

Fdef Ductile deformation e�ort. 25

Kc Fracture toughness. 23

LL Lateral load. 27

LN Normal load. 27

Mi Molar mass of the ith component. 32

Pf Probability of failure. 50

Tg Transition temperature. 30

VL Longitudal wave velocity. 30

V +R Volume ratio of pile up recovery. 39

V −R Volume ratio of sink in recovery. 39

3

Page 18: Tribological studies on the surface of glasses by lateral

4 Glossary

VR volume ratio of annealing recovery. 37

VT Transversal wave velocity. 30

V +a Pile up recovery volume after annealing. 39

V −a Sink in recovery volume after annealing. 39

V +i Initial pile up volume. 39

V −i Initial sink in volume. 39

Vi Theoritical molar volume. 32

Φ Diameter. 30

α Cutting angle of specimen. 44

β Rake angle. 25

λ Repetition distance of ripples. 44

µ Coe�cient of friction. 25

ν Poisson ratio. 30

ω The resonance frequency of the system. 47

π The ratio of a circle's circumference to its diameter. 32

ρ Density. 30

σu Threshold parameter for stress. 50

θ Apex angle. 19

fi Molar fraction of the ith component. 32

rA Shannons ionic radii of the involved ion species. 32

A Area. 46

AFM Atomic force microscopy. 24

B Brittleness parameter. 23

BMG Bulk Metallic Glass. 11

Page 19: Tribological studies on the surface of glasses by lateral

Glossary 5

COF Coe�cient of friction. 59

CSM Continuous sti�ness measurement. 32

DSC Di�erential scanning calorimetry. 30

E Young modulus. 30

EF Edge forward con�guration of tip. 53

FFT Fast fourier transformation. 44

G Shear modulus. ix, 34

H Meyer's hardness hardness. 23

h Penetration depth. 56

K Bulk modulus. ix, 34

k Lateral sti�ness. 46

m Weibull modulus. 50

N Avogadro number. 32

n Number of scratches per experiment. 50

OM Onset of microabrasion. 50

R Correlation coe�cient. 51

SEM Scanning electron microscopy. 30

SLS Soda-lime silica glass. 20

SNL Sharp Nitride Lever. 30

SSD Subsurface damage. 14

StD Standard deviation. 39

v Scan velocity. 43

x distance between two points. 46

Page 20: Tribological studies on the surface of glasses by lateral
Page 21: Tribological studies on the surface of glasses by lateral

Chapter 1

Introduction

Although some of the questions regarding the behavior of glasses under sharp

contacts have been answered in the last decades, it is still a very active topic

throughout the world's glass community, in both the academic and industrial

�elds. Considering the very outspread applications of glasses in everyday life

such as lenses, windows, monitors and touch-screen devices, surface quality

and visual appearances of employed glasses becomes an important topic. The

advances in the last 15 years in the area of more resistant glass compositions

with regard to contact-induced damage have had such prominent e�ects in

our lives that anyone with a smart phone or laptop can sense the �ourish-

ing progress. The widely used glass in ultra-sophisticated applications as

mobile devices and biomedical applications needs grinding and polishing to

the utmost precision [2�4]. Material removal in these processes is obtained

by cumulative scratching actions of multiple individual grits of random dis-

tribution, which is far from well understood [5] (e.g. frictional shear force

vector that is associated with material �ow lines in contact surface between

indenter and material [6]). On the other side, by being subjected frequently

to abrasive loads, not only the surface quality of glasses is a�ected negatively

by appearance of �aws, but also these �aws can act as stress ampli�ers and

reduce the mechanical properties of the material.

To resolve �nally the contact-induced damage issues in glasses, it is nec-

essary to take the next steps with a mechanical approach through which a

better understanding of deformation behavior of glasses under sharp con-

tact is targeted and also to determine the constitutive laws for plasticity

7

Page 22: Tribological studies on the surface of glasses by lateral

8 Introduction

and damage nucleation with respect to the structure of the glass and its

deformation under especially compressive loads. Although nanoindentation

testing has become a standard testing mean to characterize various mate-

rial properties such as elastic modulus (E), Poisson's ratio (ν), hardness (H)

and critical stress intensity factor (KIc) and all these provide good infor-

mation from the engineering point of view, they do not transcribe neither

the complexity nor the various mechanisms that take place during plastic

deformation of glasses [1,7�11]. Whence the lateral indentation technique is

used as a main method in this dissertation to extract numerous information

from the surface of di�erent types of glasses.

The scratch test is one of the fundamental techniques to evaluate the

response of material to abrasive loads under sharp contacts [12]. A scratch

is a typical surface damage that consists of visible grooves appearing as a

result of mechanical deformation [13]. The instrumented scratch test gives

access to the X-Y-Z displacements, the longitudinal one from which the

scratch velocity is computed, the lateral one (normal to the sliding direc-

tion) and the vertical displacement or penetration depth into the surface

and to the corresponding forces (Fx, Fy, Fz), the tangential force (sliding

direction), the lateral one and the applied normal force, respectively [14,15]

and from here it is possible to characterize the material in term of scratching

hardness, wear properties, fracture toughness, materials strength, scratch-

ing behaviour of coatings, adhesion and bond strength between a �lm and

substrate [4,16�23]. However, scratching is a more severe deformation mode

than indentation. Because di�erent factors such as shear-induced deforma-

tion, stick-slip reactions and frictional shear stress/adhesive force are more

likely to give rise to a fully plastic regime [4, 24].

Despite the fact that there are established protocols for scratch testing,

these protocols provide only qualitative information. Among these investi-

gations, the distinct regimes of damage in some types of glasses have been

studied [7,25]. However, A broad variety of parameters determines this phe-

nomenology, including the rate of scratching, the indenter geometry relative

to the scratching direction, the applied normal force, glass surface conditions,

environmental atmosphere and humidity, and the presence of debris or impu-

rities on the specimen surface. So far, the concrete action of these parameters

has received only very limited attention [1, 26, 27]. This is particularly the

case for the technically relevant question of compositional dependence which

Page 23: Tribological studies on the surface of glasses by lateral

Introduction 9

has been the motivation of the study in section 4.1.1. Using lateral nano-

indentation within the plastic regimes of silica, borosilicate and soda lime

silicate glasses, an insight is given to the role of densi�cation and shear �ow

in overall permanent deformation of studied glasses.

In addition, we know that the normal indentation experiments on brit-

tle materials usually ignore contributions of friction [28] or stick-slip reac-

tions [29]. This leads to fundamental di�erences in material behavior during

normal hardness testing or in lateral contact situations [30]. So far, only lim-

ited attention has been given to this topic in the �eld of glass. From here, a

special focus was set to understand the fundamental relation between stick-

slip phenomena and topography of the scratch grooves in vitreous silica and

the related results are discussed in section 4.1.2.

Surprisingly, vitreous silica with its attractive properties such as low

thermal expansion, low poisson's ratio and high free volume has not been

studied in term of tribological properties widely and almost all so far studies

focus on the phenomenology of deformation behavior of this material. This

shortage was approached by analysing the quantitative information which

can be obtained by statistical analysis of lateral force data during scratching

of vitreous silica. The results of this study are presented in section 4.2.1.

After all, even considering the valuable properties of inorganic glasses

that make them in the modern-day world widely used where transparency,

luster, and durability of glass is needed, they are not functional in some

new industrial applications that need a combination of high resistance to

wear, high yield strength, high bio-compatibility, and being at the same

time highly deformable. Metallic glasses seem to be in this direction inter-

esting candidates and have been gaining considerable attention in the last

years. For all that, their mechanical properties have been the focus of many

studies [31�37]. But the dynamical behavior of nano-scratching in these ma-

terials has not obtained much attention. For this reason, the response of

one metallic glass under lateral indentation is investigated. The results in

section 4.2.2 show the clear role of loading in the formation of shear bands

and the di�erences that exist between lateral and normal indentations.

Page 24: Tribological studies on the surface of glasses by lateral
Page 25: Tribological studies on the surface of glasses by lateral

Chapter 2

State of the art

2.1 Deformation in glasses

The juxtaposition of two words plasticity and glass may sound incongruous or

strange, since mostly glasses are considered the archetype of materials having

a brittle behavior. However, because of diverse composition, structure and

bonding nature of glasses (from metallic to covalent), not all types of glasses

are having the same kind of permanent deformation. For example, while

oxide glasses experience a brittle fracture being exposed to tensile stresses

and being far from the transition temperature, metallic glasses (bulk metallic

glasses; BMG) act in a similar way as metallic alloys, exhibiting a large

amount of plastic deformation before fracture happens when being tested

under tensile conditions [15,38].

In oxide glasses, the deformation can be the result of atmospheric cor-

rosion, as a result of devitri�cation at surface inhomogeneities or at inho-

mogeneities in the volume where the formed crystalline inclusions produce

cracks by mis�t stresses, or as a result of mechanical contact with sharp

objects. It has been recognized for some time now that the �free volume

�, which refers to the fraction of matter having a lower atomic coordination

than that in a reference material having a dense random packing of molecules

or chains of molecules and the same composition, is of key importance to un-

derstand the deformation of oxide glasses. In these free volume regions the

mechanical coupling to the surroundings is weak and this makes the inelastic

relaxations possible by local atomic rearrangements or molecular segment ro-

11

Page 26: Tribological studies on the surface of glasses by lateral

12 State of the art

tations without signi�cantly a�ecting the surroundings. On the other hand,

through the study of the kinetics of the linear viscoelastic behavior of glasses

of all types under low stress, it has been established that such local relaxation

processes are not mono-energetic but are characterized by a wide spectrum

of activation energies that correspond to a wide gradation of free volume

sites with di�erent local coordination [38].

Now, if we regard the behavior of glass under compression, it is totally

a di�erent story. Nowadays, permanent imprints left on the glass surface by

indentation are a well-known technique to study the behavior of glass under

compressive loads [15]. The �rst traces of such permanent imprints on the

surface may have been found in relation to sleek. A type of impression on

the surface that looks like a scratch rather than fractured sharp fragments.

J.W. French [39] attributed this residual imprint to the Beilby layer (for

more information on Beilby layer see [40]), or surface �ow layer, generated

by polishing, rather than to the bulk glass which has a brittle behaviour in

the nature [15].

Later Taylor announced that a diamond tip can make permanent, crack-

free indentations in glasses. Taylor in the same year performed such perma-

nent marks on the glass surface by sharp vickers or cube like shaped diamond

tool under small loads and explained them as a result of partly plastic defor-

mation [41]. Finally, Bridgman and Simon studied the mechanical properties

of oxide glasses and showed that they can be permanently compacted and

under enough high pressure some short and middle range structural rear-

rangement will occur [42]. Nowadays, scratching of glass by a hard point is

a well established method to measure di�erent characteristics of glasses as

strength, scratch hardness, wear and damage, fracture toughness, adhesion

strength and bond strength between the �lm and substrate [4, 43].

As pointed, the relationship between structural changes and loading was

studied by Bridgman and Simon who showed that for su�ciently high pres-

sure, oxide glasses experience structural rearrangements at the short- (coor-

dination number, tetrahedral to octahedral) and medium-range order (inter-

tetrahedral bond angle Si-O-Si, n-fold rings statistics). They also showed

two important pressure thresholds: the pressure to start the densi�cation

process (around 10 GPa for silica) and the pressure for saturation of densi�-

cation (near 20 GPa for silica). Between these two thresholds the permanent

Page 27: Tribological studies on the surface of glasses by lateral

2.2. Formation of cracks in contact-induced damaged surface 13

densi�cation of glass increases by increasing the pressure. By x-ray measure-

ment, almost no modi�cation of the short-range order (Si-O bond distance)

was observed in the densi�ed amorphous phase, which was attributed to an

atomic-scale mechanism leading to some sort of local folding of the glass

network upon compression [15,42].

Now it is well known that after applying load on the glass specimen, it

initially undergoes elastic deformation which has two hydrostatic and shear

component. By removal of the load, the material returns to its original

dimensions. By applying the stress beyond the yield point of the material, it

does not return to its original dimension. In this state, the hydrostatic stress

component densi�es the glass under the load, whereas the balance of shear

stress component causes the glass to undergo plastic deformation [44,45].

2.2 Formation of cracks in contact-induced dam-

aged surface

Contact-induced fracture has been studied now for more than a century.

Hertz [46] was the �rst to study the cone crack formation under loading by a

spherical indenter. These were named Hertzian cracks after him. But it was

only in the 1950s that scientists started to study the fracture mechanisms

under contact loading, trying to decipher when, how and where the cracks

initiate.

For pyramidal indenters, and more speci�cally for Vickers indenters (squared-

based pyramid), two di�erent principle cracking systems do develop in inden-

tation experiments. The �rst one is called the radial/median crack system:

median cracks are the �rst cracks to appear under the loading and parallel

to loading axis. Median cracks nucleate below the plastic deformation zone

under the load. They do need a certain load threshold in order to nucleate.

Then they propagate in a stable way as the load increases; they have a cir-

cular shape. The radial cracks are inclined toward the sliding direction and

have a curved shape and propagate perpendicular to the scratch direction

and emerge from the edge of plastic contact impression and remain close to

the surface. Lateral cracks are subsurface cracks that form during loading

at higher loads (than median and radial cracks) and propagate during un-

loading. They form beneath the deformation zone and run quasi-parallel to

Page 28: Tribological studies on the surface of glasses by lateral

14 State of the art

the sample surface. When they merge to the surface, they form chips that

may detach from the surface [25,47,48]. With mentioned studies, it became

easier to understand the deformation regimes that take place in scratching

experiments. However, scratching is in some ways a more natural aggression

of a surface, yet it is even more complex and di�cult to understand than

the indentation test. The main di�erence of scratching experiment with in-

dentation is that the indenter is slid onto the surface while a normal load is

applied, thus increasing the shear strain.

Ahn et. al. [49] stated that in scratching experiments on soda-lime silica

glass di�erent cracking systems appear: (I) the �rst regime takes place at

loads of 0-50 mN, where a hardly visible to the naked eye plastic tray (no

cracks) is left behind at the surface. Le Houérou and colleagues [1] later

called that a microductile regime, where the radial (chevron) cracks are the

�rst �aws that appear and subsurface lateral cracks form under plastic track

in a permanent groove (Fig. 2.1).

Figure 2.1: Typical scratch pattern made on the surface of the Soda lime

silica glass by a vickers penetrator (leading edge) during a monotonic loading

cycle [1].

(II) By increasing the load, shallow lateral cracks on the sides of the

plastic groove will appear, forming a chevron-like pattern. Then, as they

grow in size with increasing load, median cracks will form underneath the

groove perpendicular to the surface and parallel to the scratch direction.

This was called the the micro-cracking regime and �nally (III) microabrasive

Page 29: Tribological studies on the surface of glasses by lateral

2.3. Parameters a�ecting deformation regimes 15

regime takes place where debris form and the material is removed from the

surface [1, 49].

2.3 Parameters a�ecting deformation regimes

The most in�uential parameters a�ecting the scratching behaviour of glass

have been investigated by di�erent researchers. Glass composition, load

level, loading speed, thermal treatment history, humidity, indenter geometry

and the glass surface conditions have been mentioned as the most prominent

factors that in�uence the damage in glasses [7,9,50,51]. In the next section

some of these parameters and their in�uence on the deformation of glass will

be discussed.

2.3.1 The indenter geometry

Many investigations have shown that the development of the cracking sys-

tem depends strongly on the indenter geometry. Not only the angle, but also

the radius and the shape of indenter a�ect the critical conditions for crack

growth [26, 52�54]. It is now established that there are two basic types of

indentation fracture pattern, depending on whether the contact is essentially

elastic (�blunt�indenters) or plastic (�sharp�indenters) [55]. The blunt inden-

ter a�ects more visible scratches on the glass surface, whilst sharp indenters

grate the surface and produce �ner and more homogeneous scratches [56].

As contacts become sharper, the plastic deformation transitions toward the

shear deformation. On the other words, the blunter contacts tend to create

more densi�cation e�ect. Blunt indenters are also useful, specially at low

loads and when measuring the tribological properties of material in elastic

zone is favoured. Here by increasing the load, the response of material in the

transitional elastic/plastic zone and fully plastic zone can also be probed [57].

Lawn and Evans (1977) calculated that the critical load for fracture of

glass from a sharp indenter is 0.02 N, much less than that required for impact

of a spherical particle (for example, 0.4 mm radius particle requires 98.5

N) [58]. Veldkamp reported that in the case of ideally sharp indenters, micro-

ductile regime takes place immediately after the application of a load below a

certain semi angle of the point [52]. Schneider found that while the diamond

cone indenter with 600 tip geometry produces mostly a micro-abrasive regime

Page 30: Tribological studies on the surface of glasses by lateral

16 State of the art

for soda-lime-silica glass, the Ritz diamond indenter produces scratches in

micro-ductile regime.

Experiments with face forward or edge forward orientation of tip can

also a�ect the cracking behavior of glasses. For example, Veldkamp [52] ob-

served that the length of the median cracks obtained during scratching of a

64.9SiO2.12.3BaO.6.2K2O3.2Al2O3.1.8CaO.1.1MgO.0.55b2O3.0.2SrO.0.2CeO

glass with a face forward orientation of tip were about a factor 2 lower than

those found by Peter [59], who used a leading edge indenter in his experi-

ments.

2.3.2 The applied normal force

Load dependence of scratching behaviour of silicate glasses has been dis-

cussed by various researchers and it has been shown that the appearance of

cracks is load dependent [26, 38, 52, 53, 60�65]. Auerbach [66] showed that

there is a critical load at which the glass starts the cracking [67]. This was

further con�rmed in the experiments of Marshall [60].

The scratch pattern also strongly depends on the level of the normal load

and di�erent types of cracks appear due to the loading level (as shown in Fig.

2.1). Bensaid et al. [68] studied the e�ect of loading on the scratch pattern in

a soda lime silica glass. They found that at load of 0.1 N only median cracks

appear, from 0.3 N to 0.7 N lateral cracks initiate and propagate inside the

material without intersecting the surface, Beyond 1 N until 2 N lateral cracks

were intersecting the surface and resulted in chips growing while increasing

the load, all these in con�rmation with the observations of Veldkamp [52].

They additionally observed that the micro-abrasive phenomena was occur-

ring at the end of the scratch due to the plowing of the indenter with glass

debris (chips and fragments).

The spatial extent and the number of damage increases with increasing

the load [52]. For example, Gu [63] reported that the scratch depth and

damage zone size (It was assumed that the size of the damage zone is induced

by lateral cracks) increases in BK7 glass with increasing the applied normal

load when sliding a vickers indenter on the surface. Marshall showed that

the lateral crack length increases with increasing the normal load on the

surface of Soda-lime silica glass [60]. It has also been con�rmed that the

residual �eld provides the driving force for crack evolution and the residual

Page 31: Tribological studies on the surface of glasses by lateral

2.3. Parameters a�ecting deformation regimes 17

�eld is related to the loading. The principal stress and the residual stress

�eld beneath the indenter were studied by many researchers for glass and

ceramics [69�72].

2.3.3 Scratching rate

The results of studies performed by various researches show a relationship

between damage-induced surface and scratching speed. The type and extent

of such damage has been discussed in di�erent investigations [26,38,52,53,73�

76]. Peter [59] found that the depth of the cracks experimentally decreases

with increasing scratching speed.

Bandyopadhyay et al. [76] investigated the e�ect of scratching speed (100,

500 and 1000 µm/s) with a Rockwell C diamond indenter on the scratching

behaviour of a soda-lime-silica glass. He showed that more surface but less

sub-surface damage is induced to the glass surface at lower scratching speed

as a result of higher maximum shear stress at higher scratching speeds. He

explained this as a result of smaller contribution of shear stress just under-

neath the indenter. Furthermore, an inverse law dependency was observed

between scratch width, depth and wear volume and scratching speed for a

given applied load. That means for any applied load, the scratch width was

highest at lowest scratching speed and lowest at highest scratching speed

due to higher time of contact between the scratching indenter and the glass

surface. Also, the number of Hertzian tensile cracks and the microdamage

inside the scratch grooves reduced with scratching speed. For higher veloc-

ities, the interaction among Hertzian tensile cracks and the microdamage

inside the scratch groove were lower. Besides, At higher applied loads of

10 and 15 N micro-chip formation, as well as micro-wear debris formation

occurred and the degree of their occurrences were more signi�cant at lower

scratching speeds (e.g. 100 µm/s) than at higher scratching speeds. Another

consideration in these experiments was the relationship between shear bands

formation and scratching speed. At higher loads where the shear bands ap-

peared, they were mutually perpendicular to each other at scratching speed

of 100 µ/s, while at a higher scratching speed of 1000 µm/s the shear bands

were oriented at small angles with respect to the direction of scratching and

also with respect to each other [76].

Page 32: Tribological studies on the surface of glasses by lateral

18 State of the art

Li and his colleagues performed a series of scratches on the surface of

Soda-lime silica glasses using three conical spherical indenters (3 mm tip

radius) with 60, 90 and 136 °apex angles (2θ) under a normal load of 10

g and scratching speeds of 2, 3, 4, 5, 7, 10, 20, 50, 100 and 200 µm/s and

measured the depth and width of scratches by atomic force microscopy. They

noticed that the scratch grooves decreased with the increase of speed. For

the 90 °indenter, the crack density (de�ned as the fraction of the scratch

covered by cracks) decreased with increasing speed. This was explained by

nucleation theory: Since a crack has to be nucleated from the scratch, it may

take some time which depends on the local stress concentration based on the

usual heterogeneous nucleation theory. After a crack has been nucleated

and propagated the stress concentration is reduced so the next crack cannot

be nucleated until the stress concentration is built up again. From here by

increasing the scratching speed, the probability of cracking is reduced based

on both of these considerations. They estimated the relationship between the

crack density and scratching speed based on �rst order kinetics and showed

that the probability of cracking is proportional to the available uncracked

area of the scratch and the faster indenter moves, the bigger becomes the

uncracked area.

This relationship was also observed by Klecka et al. [74] in the way that

the faster scratch speed contributed to the increased propagation of lateral

cracking around the edges of the scratch, resulting in further material re-

moval. Both studies con�rm the observations of Veldkamp (1978) [52] that

number of cracks per unit length of scratch decrease with increasing the

speed of sliding. However, it has been shown that when the crack density

was low or when the separation of cracks exceeded a certain distance, the

interaction between cracks no longer played a role so the density of cracks

became independent of scratching speed or it was assumed that it depended

on speed in a di�erent way. The coe�cient of friction (the horizontal to

vertical force ratio) was shown to be independent of scratching speed in the

range of their experiments [73].

2.3.4 The glass surface condition

The topic of glass surface is very wide and consists of chemical and physical

reaction of environment with glass, surface tensions and relaxation, topog-

Page 33: Tribological studies on the surface of glasses by lateral

2.3. Parameters a�ecting deformation regimes 19

raphy, the presence of debris, particles on the surface from polishing or pro-

duced by applying load on the surface and so on. In fact, the glass surface

condition could be itself a independent topic of investigation. But discussing

all various aspects of glass surface is not within the scope of this work and

hereafter, only some facets that are in relation to our investigations will be

considered here.

Any material interacts with its environment through its exposed surface.

The physical and chemistry status of a bulk glass is not necessarily trans-

ferable to the bulk glass and hence, characterization of the glass surface

presents unique challenges and opportunities [77�79]. The manufacturing

process can a�ect the structure of glass surface dramatically. Surface �aws

produced by machining or handling are the predominant source of failure in

glass [80]. From here, the engineering of glass surfaces could also provide

the next breakthroughs in glass technology, for example, in the design of

glasses with special surface properties such as high chemical durability, high

damage resistance, self-cleaning properties, or other types of functionalized

surfaces. Nearly all oxide glass surfaces are almost immediately hydroxy-

lated by exposure to a humid environment which was explained in section

2.3.5 [79].

The presence of impurities, debris, microscopic cracks makes the glass

more vulnerable to unwanted deformation. The microabrasive phenomena

occurring at the scratching process is due to the plowing of indenter with

glass debris and the chip formation is reported to be as a result of shear-

induced microcracks generated in the sub-surface region of some glasses (for

example SLS) (chips and fragments). Thus, the polishing processes and

materials can also have a strong in�uence on the surface quality of glasses.

Veldkamp [52] observed more lateral cracks on the mechanically polished

surface of glass rather than �ame polished one [7, 27, 63, 68, 76]. When such

microchips fall in between the sliding indenter and the scratch groove, they

are further comminuted to form the micro wear debris which get entrapped

in between the sliding indenter and the scratch groove, and enhance the

coe�cient of friction [44,81].

Page 34: Tribological studies on the surface of glasses by lateral

20 State of the art

2.3.5 Environmental atmosphere and humidity

Another parameter a�ecting the deformation regimes of glass is humidity.

It is believed that humidity can a�ect the glass through a stress-enhanced

chemical reaction between a chemical environment, usually water, and the

Si-O-Si bonds. From here, it has been reported that the water environment

promotes the crack initiation in silica, pyrex type glass and soda-lime glass

[82, 83]. Since our studies focus mostly on silicate glasses, the following

description will be about SiO2 network. However, there are some studies

about other glass compositions such as phosphate glasses [84], borosilicate

glasses [85], and germanate glasses [86]. In the vicinity of crack tip, the

silica network can be deformed. This deformation mainly impacts Si-O-Si

inter-tetrahedral bond angle (The O-Si-O tetrahedral bond angle will also be

a�ected, but to a much lesser extent). Hence the polar moment of Si-O bond

will be modi�ed through deformation and this is very important in term of

energy that is required for H2O molecules to approach Si-O-Si bonds [87�89].

Michalske et al. [87] proposed a three step mechanism for the hydrolysis of

silica network:

1. The glass surface adsorbs the water molecule

2. Two hydrogen bonds are created, one between the silicon atom be-

longing to the Si-O-Si bridge and the oxygen atom from the water

molecule, and the other between the oxygen atom belonging to the

siloxane bridge and the hydrogen atom from the water molecule

3. The Si-O bond is broken, and two silanols (-Si-OH function) are cre-

ated.

SiO2 +H2O ↔ 2Si+OH. (2.1)

It was revealed that a decrease in the inter-tetrahedral angle (Si-O-Si),

coupled with a slight decrease in the tetrahedral angle O-Si-O, results in a

tremendous increase in the reaction rate. It should be mentioned that it is

not the increase in the Si-O-Si bond angle that causes the lowering of the

total energy barrier for hydrolysis to occur, but instead it is a decrease in

this bond angle that may occur through a local pinching of the glass network

Page 35: Tribological studies on the surface of glasses by lateral

2.3. Parameters a�ecting deformation regimes 21

that may exist locally within edge-sharing structures or resulting from the

deformation of the 3-D glass network as a consequence of the high strain

gradient that exists in the vicinity of a loaded crack tip. An increase in

the dissolution rate of the silica and soda-lime silica glass was seen for the

densi�ed glass in indentation experiments [90].

The humidity can in�uence the appearance of types of cracks as well.

While there is almost a total lack of radial cracks at 0% humidity, the lateral

cracks, which never reach the surface, extend over a large distance in soda-

lime silica glass. However, Sub-surface lateral cracks, formed just behind

the indenter, were found to be insensitive to the hygrometric rate and to

remain under the surface of the material [1]. Furthermore, large chips can be

observed at 30% relative humidity. In contrast, the grinding regime is barely

reached at 3 N in a humid environment, whereas at 1 N it is well developed for

dry conditions. Both friction coe�cient and subcritical crack growth, which

are humidity-dependent, do play a major role here [54]. The humidity has

been reported to decrease the strength of glass as well. Proctor et. al. [91]

explained that the strength of silica degraded by increasing the humidity of

environment and this was attributed to the rate of reaction of water molecules

at the stressed crack tip, the mobility of water molecules toward the crack

tip and the availability of water molecules. The humidity causes on one side

the transition between di�erent regimes to develop earlier in the scratching

experiments and on the other side increases the crack velocity [1, 15, 92, 93].

If the crack size is reduced su�ciently, the water can always di�use to the

crack tip at a rate su�cient to provide a stress corrosion environment.

Doing indentation experiments in other environments like bu�er solutions

has also been reported. Mould found that the strength of soda-lime glass

was relatively insensitive to bu�er solutions ranging in pH from ∼1 to 13.

He found that the strength increased for pH values greater than 13 and

decreased for pH less than 1 [80, 94]. In summary, it can be said that the

water in glass is known to decrease the mean connectivity of the network,

thus leading to a decrease in properties such as Young'modulus and viscosity

as its content rises [95].

Page 36: Tribological studies on the surface of glasses by lateral

22 State of the art

2.4 Compositional dependence of deformation in

glass

As with most materials, the mechanism of inelastic deformation of glasses is

closely linked to their structure. When exposed to local mechanical contact

glass surfaces undergo permanent microscopic deformation reactions [41,80].

The type and extent of these reactions is strongly composition-dependent and

have been studied by various researchers [96�99]. It is now widely accepted

that in �free volume�regions, where mechanical coupling to the surroundings

is weak, inelastic relaxations become possible by local atom rearrangements

or molecular segment rotations without signi�cantly a�ecting the surround-

ings [80].

Using normal indentation experiments, a general distinction has been

made between �normal�and �anomalous�glasses according to the presence of

shear �ow and structural compaction, respectively [100�102]. It was found

that the degree of possible structural compaction correlates with the atomic

packing density and the Poisson's ratio of any given glass composition, pro-

viding parameters and a guideline for dedicated chemical tailoring [103].

Based on this understanding, less brittle glasses have been discovered [104]

which o�er increased defect resistance. A large number of studies have been

following this route [105]. Lawn and Marshall [106] introduced a brittleness

parameter B, that accounts for two competitive processes, plastic deforma-

tion and crack propagation and it is calculated as:

B = H/Kc (2.2)

Where Kc is the fracture toughness and H is the Meyer's hardness. The

higher the B, the more brittle is a material and from here it is possible to

compare the brittleness of the glasses of the same family.

The contributions of shear �ow and structural compaction to the over-

all of permanent deformation can be quanti�ed through classical relaxation

experiments [107]: using a normal indentation test, volume and topography

of the residual imprint are compared to those of the same imprint after ex-

posure to a temperature around Tg for prolonged time [108,109]. Assuming

full recovery of structural compaction as a result of thermal relaxation, the

degree of compaction is then obtained from the volume di�erence while the

Page 37: Tribological studies on the surface of glasses by lateral

2.4. Compositional dependence of deformation in glass 23

residual is related to shear �ow. Studies of this type have been conducted

on a variety of glass types, e.g., Refs. [108�112].

Here, coordination number evolution can play a role too. Fused silica

and window glass start to permanently densify above 10 GPa, then reach a

saturation level above a pressure of 20 GPa, whereas silica has a saturation

level of permanent densi�cation ratio of 21%. Soda-lime silica glass, such as

a window glass, saturates at a 6.5% densi�cation ratio (Fig. 2.2).

Figure 2.2: Maximum density relative variation after high hydrostatic pres-

sure testing at room temperature as a function of poisson's ratio. The atomic

packing density Cg for some compositions is also given [103]. Reprinted

from [15]

This saturation level was shown to be linked in some way to the Poisson's

ratio of the glass [103, 113, 114] and Poisson's ratio is correlated to atomic

packing density, Cg. Cg is de�ned as proportion of minimum theoretical

volume occupied by the atoms to the corresponding e�ective volume of glass.

The glasses with low atomic packing density deform mostly by densi�-

cation mechanism and the ones with closed pack systems go mostly through

shear deformation mechanism. In general, the densi�cation process is asso-

ciated with a signi�cant decrease of the intertetrahedral angle, for example

Page 38: Tribological studies on the surface of glasses by lateral

24 State of the art

Ge-Se-Ge in chalcogenide glass GeSe4 and Si-O-Si angles in SiO2 glass. This

can be due to small or undetectable changes of the interatomic distances

(adjacent neighbors), and a gradual increase of the average atomic coordi-

nation (from 4 to 6 in amorphous silica, from 3 to 4 in B2O3 glass, from

fourfold to sixfold for Ge and from twofold to fourfold for Se in GexSe1-xglasses) [101,115].

The transition from normal to anomalous behaviour by adding sodium

and calcium oxides to SiO2 was shown clearly by Limbach and his col-

leagues [112]. For anomalous glasses such as silica, glass deforms essentially

by densi�cation, while in normal glasses the plasticity composed of shear

�ow [100]. Le Houérou and his colleagues showed that the scratchability is

greatly a�ected by composition moving from silica toward di�erent composi-

tions of soda-lime-silica (SLS) glasses. For example, in ramp load scratching

experiments, the microabrasive regime appears in low-load domain for fused

silica and SLS glasses with low silica contents, whereas the lateral chipping

happens for samples with higher amount of silica. Higher amount of silica

leads to earlier appearance of microabrasive regime, except for fused silica

which has an anomalous behaviour.

In summary, it seems that glasses from the devitrite phase �eld that

include crystal nucleations, are sensitive to chipping and glasses with silica-

like networks appear to be much more resistant to both crack propagation

and chipping during scratch experiments. At atomic or molecular scale,

the glasses with high silica content have a better resistance to chipping and

micro-cracking and that is due to the open structure of these glasses network

that allows for both network �exibility and �ow-densi�cation [1, 15,97].

2.5 Contribution of friction to materials behaviour

in indentation experiments

Although there are fair number of studies on the frictional sliding behavior of

various materials such as nanocrystalline ferrite [116], metals and lubricant

oils [117], contact lenses [118], stainless steel sheets [119], graphene [120],

polymers [121�124], aluminum and steel tools [125], the number of such

investigations on glasses under sharp contacts is limited.

Page 39: Tribological studies on the surface of glasses by lateral

2.5. Contribution of friction to materials behaviour in indentationexperiments 25

The friction coe�cient is a critical parameter in designing the mechan-

ical systems with contacting surfaces, modelling of friction-induced vibra-

tions, mechanism of lubrication and where the comfort of soft movements

are needed. However, modelling frictional behaviour is not simple, since

friction force depends on various parameters such as surface roughness, true

contact area, normal load, dynamic behaviour of contact interface with vi-

bration, material transfer, sample thickness, test con�gurations and sliding

systems [4, 119,126�128].

The �rst systematic study of the friction of glass was reported by the

Hardys [129, 130] in 1919, who studied the spreading of �uids on glass [67].

Since then investigations have been carried out to understand the relation-

ship between di�erent parameters such as surface condition, loading, scratch-

ing speed and humidity with friction coe�cient [68]. For instance, a clean

surface without any debris or thin �lms has a high friction coe�cient. The

presence of humidity at the interface results in liquid-mediated adhesion,

which may result in higher friction. It is also being reported that the friction

coe�cient increases as a function of load almost linearly for glass iron, and

aluminum sliding on glass in vacuum and atmosphere saturated with H2O.

However, this behavior is not the same in all conditions. For example, in

the case of glass sliding on glass, the coe�cient of friction remained unaf-

fected by increasing the pressure from 10-10 torr to 1 torr, but from 1 torr

to atmospheric pressure, the coe�cient of friction increased. In term of en-

vironmental conditions, the adhesion force for glass in the presence of water

is more than three times that for glass in the presence of octane [131,132].

Bensaid and his colleagues [68] investigated the relationship between fric-

tion coe�cient and loading for a soda lime silica glass by ramp load scratch-

ing. They introduced a model to calculate the approximate damage loads in

scratching experiments by considering the cracking threshold. In the case of

scratching of the glass, they considered that the tangential e�ort (Ft) should

be divided to a ductile deformation e�ort (Fdef), a coulumbian friction e�ort

(Ff), and a cracking e�ort (Fcr):

Ft = Fdef + Ff + Fcr (2.3)

Page 40: Tribological studies on the surface of glasses by lateral

26 State of the art

and thus,

µ0 =Ft

W=

Fdef + Ff + Fcr

W= µdef + µf + µcr (2.4)

where µdef = tan(β)= constant (considering a pure plastic deformation),

and β is the rake angle, µf =Ff

W = constant (considering a Coulombian

friction).

Assuming Fdef and Ff constant, they considered the apparent friction of

coe�cient to be a result of the cracking e�ort and hence dependent to the

depth of cracking. By setting the depth of crack to zero, a threshold was

found for loads below which there is no median cracks and over which the

transition load for radial/lateral cracking will start. They further found that

the friction coe�cient increases with the increase of the load until a certain

value and then beyond this load it stabilizes. Liu [133] showed that the

friction coe�cient �rst decreases and then increases with increasing normal

load when conical spherical indenter slides over a fused silica glass, while Wei

[134] and Liu [?] revealed that coe�cient of friction has di�erent behaviors

when increasing the load at various scratching velocity on the same material.

One fact is clear: friction coe�cient is not only a function of load, but also

a proper loading range should be chosen to observe any variation of it.

Moreover, the friction of coe�cient is also velocity rate dependent. One

major consequence of �uctuation of friction force or sliding velocity is the

creation of some oscillations that are called �stick-slip�phenomenon. The

term �stick-slip�was coined by Bowden and Leben [135] and can degrade the

materials performance signi�cantly by causing vibrations that lead to cracks

and wear [136]. During the stick phase, the friction force builds up to a

certain value and once a large enough force has been applied to overcome

the static friction force, slip occurs at the interface [44]. The latter phe-

nomenon can be accurately investigated in a variety of controlled conditions

(e.g., loading, scan velocity, and temperature) using atomic force microscopy

(AFM) [137�140].

As an example, it has been recently demonstrated that the formation

of nanoripples on polymers can be understood as a combined e�ect of the

time-increasing shear stress imposed by the slider (i.e., the sharp probing

tip) and the viscoplastic response of the material. In this way, the rippling

Page 41: Tribological studies on the surface of glasses by lateral

2.5. Contribution of friction to materials behaviour in indentationexperiments 27

process appears to be ruled by �ve parameters: normal force, indentation

rate, scan velocity, tip geometry, and lateral contact sti�ness [29,141�143].

Page 42: Tribological studies on the surface of glasses by lateral
Page 43: Tribological studies on the surface of glasses by lateral

Chapter 3

Experimental methodology

3.1 Ramp load scratching

By ramp load scratching the response of material can be studied over a range

of loads in a single test rather than many constant load tests and usually

a transition from elastic to plastic deformation on the surface is observed.

In this test, a tip is brought into contact with the sample, Then the tip is

loaded with a constant loading rate, while simultaneously translating the

sample. Prior to and after the scratch test, a single line scan of the surface

topography is completed for comparing the original surface to the deformed

one. Therefore, each ramp load scratching tests comprises of three steps:

a single-line prescan of the area to be tested, the ramp load scratch test,

and a �nal scan to evaluate the residual deformation [144]. The deforma-

tion regimes induced by ramp load scratching were studied in a commercial

vitreous silica and a metallic glass sample.

3.1.1 Ramp load scratching test on vitreous silica

Instrumented nanoindentation (G200, Agilent) was employed to generate

quantitative data on the scratch resistance of commercial grade vitreous

silica (Heraeus Suprasil 1). The experiment comprises control of the normal

load LN on a Berkovich tip and recording of the lateral load LL during lateral

displacement as illustrated in Fig. 3.1.

29

Page 44: Tribological studies on the surface of glasses by lateral

30 Experimental methodology

Figure 3.1: (a) Schematic of the determination of lateral load LL during

lateral displacement of Berkovich indenter

The value of LL results from a speci�c rate of normal loading and lat-

eral displacement. It is determined from the lateral sti�ness of the indenter,

KL, and its displacement in x- and y-directions, shown schematically in Fig.

3.1. The overall observation length (lateral displacement) was kept constant

among all samples (1.0 mm). Samples themselves were cylindrical with a

diameter of 33 mm and thickness of approximately 3 mm. On the studied

surface, they were sequentially polished with dry silicon carbide powder with

grain sizes of 70, 40 and 9 µm, and �nally with a suspension of diamond pow-

der with a grain size of 1.0 µm, leading to an average roughness of 1.19 µm

(mean arithmetic height, taken from confocal microscopy) and subsequently

stored in vacuum.

Directly before analysis, the samples were cleaned in an ultrasonic bath

of pure isopropanol for 5 min at room temperature, and subsequently �ushed

with ethanol. Tests were conducted by increasing the normal load LN from

0.05 mN to 300 mN during lateral displacement at rates of 10, 50, 100, 150,

300 and 500 µm/s across the overall lateral displacement range of 1.0 mm,

at room temperature. This corresponds to normal loading rates between 3

mN/s and 150 mN/s. The employed tip geometry is shown in Fig. 3.2.

Scanning was conducted in edge forward con�guration (EF, Fig. 3.3).

For each test, an initial specimen surface pro�le was obtained before scratch-

ing by pre-scanning the sample's surface with the indenter under a load of

Page 45: Tribological studies on the surface of glasses by lateral

3.1. Ramp load scratching 31

Figure 3.2: a 3D representation of the employed tip, obtained by wide-�eld

confocal microscopy.

50 µN. While testing, both the penetration depth and the value of LL were

continuously monitored. After scratching, the surface pro�le of the sample

was scanned again under the same conditions as during the pre-scanning

stage. For each loading rate, 20 scratches were performed.

Figure 3.3: The tip con�gurations during scanning, edge-forward (EF, ap-

plied here) and face-forward (FF)

Page 46: Tribological studies on the surface of glasses by lateral

32 Experimental methodology

3.1.2 Scratching of metallic glass by a spherical indenter

Indentations and scratch tests were performed on the surface of Zr55Cu30Al10Ni5metallic glass. The samples were received from France and a thorough de-

scription of sample preparation can be found in ref. [145]: They were pre-

pared using an arc melting method with a mixture of pure Zr, Cu and Al met-

als in an argon atmosphere. By using an arc tilt casting method, cast BMG

cylindrical rod specimens (Φ: 8 mm × L: 60 mm) were obtained [145�147].

A high speed diamond/copper saw was used in order to get 50 mm long sam-

ples. The density, ρ = 6.776 g/cm3 was measured using Archimedes'principle

technique. The Young's modulus E = 85 Gpa and Poisson's ratio ν = 0.370

at room temperature were measured by ultrasonic echography using 10 MHz

piezoelectric transducers in contact with the sample via a coupling gel. E

and ν are expressed as:

E = ρ(3V 2L − 4V 2

T )/((VL/VT )− 12) (3.1)

and

ν = (3V 2L − 4V 2

T )/(2(V2L − V 2

T )− 1) (3.2)

where VL and VT are the longitudinal and transversal wave velocities, re-

spectively. The Tg = 688 K was measured by di�erential scanning calorime-

try (DSC) [148].

In Otto Schott Institute of Materials Research a conical diamond tip (E

= 1141 GPa, Poisson's ratio = 0.07, e�ective tip radius = 4690 nm) was used

to perform indentation imprints at loads of 10, 20, 30, 40, and 50 mN. The

scratching experiments were performed at the same indentation loads and on

the lengths of 100, 200, 300, 400 and 500 µm subsequently. A loading rate

of 1 m/N was chosen for all scratching experiments. After a pre-scanning of

the surface with the load of 50 µN, the experimental load was applied and

in the end of it the scratch was scanned again with a load of 50 µN. Both

lateral load and penetration depths were recorded during this phase. For

each experimental condition, 5 tests were carried out. A scanning electron

microscope from JEOL company (JSM 7001F) was used to investigate the

sample surface after indentations and scratchings.

Page 47: Tribological studies on the surface of glasses by lateral

3.2. Constant low load scratching 33

The surface was further investigated by an atomic force microscope (AFM)

device as well. The measurement were carried out in Leibniz Institute of Pho-

tonic Technology (IPHT) in contact mode using a SNL tip A (Sharp Nitride

Lever) from Bruker. The device was the model Dimension Edge, equipped

with a 100 µm × 100 µm scanner and a maximum z-range of 10 µm. The

AFM data were analysed by WsxM 4.0 [149] software.

3.2 Constant low load scratching

In constant load scratch testing, the normal force is maintained at a con-

stant level while scratching the sample. By this method, the friction can

be measured during scratching [44]. The scratching test at constant load is

performed for a variety of reasons: to characterize material, to determine

the e�ects of variables, and to select the material for a speci�c application.

Two studies were performed in the constant load scratching: one on vitreous

silica and the other on three di�erent silicate glasses.

3.2.1 Scratching at constant load by a Berkovich indenter on

vitreous silica

The scratch tests and AFM imaging were performed on commercial-grade

vitreous silica (Heraues Suprasil 1). A load-controlled nanoindenter (G200,

Agilent) equipped with a Berkovich diamond tip was used to generate surface

scratches. After a pre-scanning phase with a loading force below 50 µN,

the tip was suddenly halted and the load was increased up to the chosen

setpoint (between 10 and 30 mN, i.e., within the elastic-plastic regime [150]).

Scanning was resumed shortly after until the desired scratch length was

reached. At this point, the tip was again halted, and the loading force

brought back to the prescan value. Together with the penetration depth,

the lateral force was recorded during the whole process.

The resulting scratch grooves were subsequently imaged using AFM (Nano

Wizard 4, JPK Instruments) in contact mode in the group of professor Enrico

Gnecco. Silicon nitride probes (Tip B, SNL-10, Bruker) with force constant

of 0.12 N/m and resonance frequency of 23 kHz were used, and the fast scan

direction was oriented along the scratch direction. All images were analysed

using JPK proprietary software and WSxM 4.0 [149].

Page 48: Tribological studies on the surface of glasses by lateral

34 Experimental methodology

3.2.2 Constant low load scratching on silicate glasses

3.2.2.1 Glass samples

Commercial-grade vitreous silica (Suprasil 1, Heraeus), a borosilicate glass

(Boro�oat BF33, Schott), and a standard soda lime silicate glass (Optiwhite,

Marienfeld) were chosen for this study. Compositions of these glasses are

provided in Table 3.1. The silica samples were polished stepwisely on dry

silicon carbide powder with grain sizes of 70 µm, 40 µm and 9 µm. In the

�nal polishing step, a suspension of diamond power with a grain size of 1

µm was used. The other two glass types were used as received. Specimens

of about 30 mm in diameter and 2 mm (silica and soda lime silicate) or 1.1

mm (borosilicate) in height were cut from the glasses and stored in vacuum

between experiments. Directly before indentation studies, the samples were

rinsed with acetone and dried in �owing nitrogen.

3.2.2.2 General characterization

Physical properties of the three glass types are provided in Table 3.1. They

were obtained for this study by instrumented indentation using a nanoin-

denter platform (Agilent G200) and ultrasonic resonance, respectively [151].

Table 3.1: Glass transition temperature Tg, density ρ, Young's modulus E,

shear modulus G, bulk modulus K, atomic packing density Cg and Poission's

ratio ν of the studied glasses.

sample name Silica SLG BF33

Composition SiO2(Suprasil 1)72SiO2.13.9Na2O.8.8CaO.4.3MgO.

0.6Al2O3.0.4K2O.0.2SO3.0.02Fe2O3

81SiO2.13B2O3.4Na2O/K2O.2Al2O3

Tg(0C) 1120 545 525

ρ(g/cm3) 2.203 2.569 2.215

E(GPa) 72.0 71.0 62.3

G(GPa) 31.2 28.8 26.1

K(GPa) 34.6 44.5 33.7

Cg 0.457 0.514 0.475

ν 0.153 0.234 0.192

All indentation and subsequent scratching experiments were performed

in air at 25 0C. Normal indentation was done at constant strain-rate in

the continuous sti�ness measurement (CSM) mode, allowing for continuous

Page 49: Tribological studies on the surface of glasses by lateral

3.2. Constant low load scratching 35

measurement of sti�ness by superimposing a small oscillation on the primary

loading signal and analysing the resulting response of the system. The den-

sities ρ were determined with the Archimedes method at 25 �in ethanol.

From this, the packing density Cg was estimated, [152].

Cg = ρ

∑︁fiVi∑︁fiMi

(3.3)

where Vi = 4/3πN (xrA3 + yrB3) is the theoretical molar volume of the

ions of a generic compound AxBy, Mi denotes the molar mass of the ith

component present in the molar fraction fi, N is the Avogadro number, and

rA and rB are Shannon's ionic radii of the involved ion species (using rO =

135 pm) [153].

3.2.2.3 Scratching and indentation tests

Since literature data on indentation volume recovery by thermal relaxation

exhibit notoriously high scatter [108, 109], we initially conducted original

reference experiments by normal indentation. For comparability with the

subsequent lateral tests, these were done with a 3-sided Berkovich tip (in-

stead of a Vickers tip which is used in most of the literature studies). Lateral

testing was subsequently performed by using the tip in edge-forward con�g-

uration with constant normal loads of 30, 23, and 17 mN on silica, SLG and

BF33, respectively, and two di�erent loading rates for each sample (10 and

50 µm/s). The loads were chosen so as to avoid the occurrence of chipping or

microabrasion. In each scratching experiment, a pre-scan was conducted at

a low loading force of 50 µN. After that, the desired load was applied rapidly

(limited only by the indenter response rate) and kept constant on lines of 35

µm (BF33 and SLG) and 70 µm (silica), respectively. A post-scanning phase

with the same load value as the pre-scan concluded each scratching measure-

ment. All experiments were carried-out at ambient humidity, using the same

tips and shortest possible delay times (<12 h) between each experiment.

3.2.2.4 AFM imaging and subsequent heat treatment

After normal and lateral indentation, the residual indents and the scratch

grooves were imaged by atomic force microscopy (AFM, Nano Wizard 4, JPK

Page 50: Tribological studies on the surface of glasses by lateral

36 Experimental methodology

instruments) in tapping mode in the group of professor Enrico Gnecco. Prior

to that, the samples were again cleaned in an ultrasonic bath of ethanol for

5 min at room temperature, �ushed with acetone and dried properly. Silicon

probes (PPP-NChauD, Nanosensors TM) with a nominal force constant of 42

N/m and a nominal resonance frequency of 330 kHz were used for mapping.

Thermal annealing was conducted in dedicated furnaces with high tem-

perature stability. For SLG and BF33, we used a dilatometer furnace (Net-

zsch) with �ne temperature control and a �at pro�le over a length of <5

mm. The silica sample was relaxed in a bottom lift furnace (Nabertherm

P 310) with an additional thermocouple positioned directly on the sample

for manual control of the local temperature. For all samples, annealing was

done at 0.95Tg, i.e., 511±2 0C for SLG, 498±2 0C for BF33, and 1060±5 0C

for silica (the Tg values were extracted from the data sheets of the glasses;

they are provided in Table 3.1). For SLG and BF33, annealing was done for

1h. For silica, two separate experiments were conducted at 1h and at 2h,

respectively.

AFM images were taken before and after annealing. The time between

images was limited only by the cooling rate of the annealing furnace; it

was <3h for SLG and BF33, and <6h for silica. Through this rapid data

acquisition, secondary relaxation reactions were excluded as far as possible.

Page 51: Tribological studies on the surface of glasses by lateral

Chapter 4

Results and discussions

Scratch-induced surface of Silica under ramp load

scratching

The optical analyses performed on the surface of vitreous silica after ramp

load scratching con�rmed the general phenomenon of sequential plastic de-

formation, chipping and micro-abrasion Fig. 4.1 [7]. Beyond the plastic

regime, the cracks which occur around the scratch groove are mostly surface

chips or lateral cracks.

The types of cracks occurring on the surface are clearly load dependent.

Also the scratch pattern is strongly dependent to the normal load level and

by increasing the load, di�erent regimes of deformation take place that are

schematically shown in Fig. 4.1.

A broad variety of parameters determines this phenomenology, includ-

ing the rate of scratching, the indenter geometry relative to the scratching

direction, the applied normal force, glass surface conditions, environmental

atmosphere and humidity, and the presence of debris or impurities on the

specimen surface, as discussed in section 2.3. So far, the concrete action of

these parameters has received only very limited attention [1, 26, 27]. This is

particularly the case for the technically relevant question of compositional

dependence. Here, recent approaches to compositional development rely

largely on the assumption that data obtained from normal indentation cor-

relate directly (or even linearly) with damage resistance under lateral contact

load. The consensus is that the indentation response of glasses is governed by

37

Page 52: Tribological studies on the surface of glasses by lateral

38 Results and discussions

Figure 4.1: (a) Typical scratch pattern which is observed on silicate glasses

during steady scratching with increasing normal load. (b) Scratch pattern

of vitreous silica observed at ramping load of 0.05-300 mN and scratching

rate of 100 µm/s.

the interplay of elastic deformation, structural compaction and shear [154].

Relaxation studies can subsequently be used to evaluate individual contri-

butions of the latter two [108, 112, 155]. Then, the ability of the considered

material to compact depends directly on its free volume, on molecular scale,

and correlates with Poisson's ratio [103]. Accordingly, vitreous silica, with

exceptionally low Poisson's ratio and high free volume, exhibits a degree of

structural compressibility which beats that of almost all other glasses. How-

ever, it has also become clear that the structural reactions which underlie

damage in�iction are signi�cantly more complex [151].

Although there are some studies investigating the relationship between

materials composition and damage resistance through indentation experi-

ments [108, 109], not much work has been done on the role of densi�cation

and shear �ow in lateral indentations. This was the motivation to study the

role of densi�cation and shear �ow in low load scratching regime (plastic

regime region in Fig. 4.1) on two anomalous and one normal glass. Addi-

tionally, some wavy patterns that appear in this region were investigated by

Page 53: Tribological studies on the surface of glasses by lateral

4.1. Studies at low load scratching in plastic regime of deformation 39

AFM. The results are presented in sections 4.1. By increasing the load in

lateral indentation, other deformation regimes take place, as mentioned and

shown in Fig. 4.1. From here, two studies were performed in higher loads on

vitreous silica and a metallic glass and the results are presented in section

4.2.

4.1 Studies at low load scratching in plastic regime

of deformation

As described in previous section, the initial ramp load scratching experiment

performed on the surface of vitreous silica provided valuable information

about di�erent deformation regimes taking place within the groove. From

here, four di�erent studies were designed that focus on low load and higher

loads (ramp load) scratching to determine some tribological properties of

three oxide and one metallic glass. Two studies which were performed in

constant low load scratching in the section of plastic regime deformation are

explained further in the next two sections (4.1.1 and 4.1.2).

4.1.1 Relaxation of scratch-induced surface deformation in

silicate glasses

In this study, the contributions of shear �ow and structural compaction to

the overall of permanent deformation was quanti�ed through classical relax-

ation experiments within the plastic regimes of silica, borosilicate and soda

lime silicate glasses. By using normal and indentation test, volume and to-

pography of the residual imprints are compared to those of the same imprint

after exposure to a temperature around Tg for prolonged time. Yoshida's

protocol of deformation volume analysis is adopted [108,109] before and af-

ter thermal treatment in order to reveal the roles of compaction and shear

�ow in comparison to normal indentation.

Data analysis The AFM images were �rst �attened using the WsxM

software [149] to exclude the slope of the untreated substrate from the anal-

ysis. The same software was used to measure the indentation volumes and

to extract the scratch groove cross-section pro�les. An exemplary image of

a typical scratch groove (BF33, after relaxation) is provided in Fig. 4.2.

Page 54: Tribological studies on the surface of glasses by lateral

40 Results and discussions

Figure 4.2: Typical AFM micrograph of a scratch groove as used for volume

analysis (here: sample BF33, scratching at normal load FN = 23 mN, 50

µm/s, after annealing for 1h at 498±2 0C).

For each scratch, the average of 10 cross-sections spaced about 2 µm

from each other perpendicular to the scratch direction was used in the fur-

ther analysis. The pile-up and sink-in areas were estimated from the integrals

of the projected cross-sections and multiplied by the length of the consid-

ered scratch section to obtain the corresponding groove volumes. In the

following, V+ and V- will denote the pile-up volume (above the �at surface)

and the sink-in volume (below the �at surface), respectively. Parameters

obtained before and after annealing have the additional subscript �i�(initial)

and �a�(after), respectively. The volume ratios VR of annealing recovery

were calculated following Yoshida et al. [108],

VR =(V −

i − V −a ) + (V +

a − V +i )

V −i

= VR + V +R (4.1)

In the 3D overview AFM image of the scratch groove of BF33 (Fig.

4.2), the pile-up region which formed after 1h of heat treatment can be

clearly seen (all three types of glasses showed some pile-up on the sides of

scratch groove after annealing below Tg). The measured volumes before

and after annealing are presented in Table 4.1 for normal indentation and

scratching. The standard deviation StD stated in the table was calculated

Page 55: Tribological studies on the surface of glasses by lateral

4.1. Studies at low load scratching in plastic regime of deformation 41

for the recovery ratio measured across 10 cross-sections alongside the scratch

groove for each scratch. It is below 10% for all samples.

As already noted, the overall shape of the scratch region is a�ected by

densi�cation (due to structural compaction of the material on molecular

scale) and shear �ow (generating a variation in body shape without a vol-

ume change, most visible in the pile-up zone) [102]. On �rst inspection,

both in scratching and in normal indentation deformation, all glasses exhibit

major but di�ering amounts of densi�cation. Although we used a Berkovich

indenter, this is in general agreement with other observations of the normal

indentation behavior [108,109].

Figure 4.3: AFM top-views and cross-pro�le scans (indicated in the AFM

micrographs) of the residual imprints after Berkovich indentation, before

(left) and after (right) annealing for 1h at 0.95Tg, shown for (a) silica, (b)

BF33, and (c) SLG.

Fig. 4.3 shows AFM top view images of Berkovich indents before and

after annealing for the three silicate glasses. Together with the data provided

in Table 4.1, silica exhibits the highest recovery ratio on normal indentation

(∼ 80%), while the SLG shows the lowest value (∼ 47%). These numbers

Page 56: Tribological studies on the surface of glasses by lateral

42 Results and discussions

Table

4.1:Indentation

andscratch

volumes

before

andafter

annealingfor

di�erent

experim

entalconditions.

Sample

name

Experim

entalconditions

Before

annealingAfter

annealingRecovery

ratioStD

V−i(µm

3)V

+i(µm

3)V

−RStD

V−a(µm

3)V

+a(µm

3)V

+RStD

SilicaScratch

at10

µm/s-1h

18.268.11

4.312.27

0.40.21

Scratchat

50µm/s-1h

17.268.2

3.832.13

0.420.11

Scratchat

10µm/s-2h

20.488.22

0.760.01

4.898.33

0.0090.11

0.770.12

Scratchat

50µm/s-2h

21.274.07

0.780.02

4.735.88

0.080.07

0.780.02

Indent-1h9.31

--

1.65-

-0.83

0.05

7.86-

-1.76

--

0.780.05

7.01-

-1.38

--

0.800.05

Indent-2h2.68

--

1.15-

-0.73

0.05

SLG

Scratchat

10µm/s-1h

7.032.04

0.460.04

3.802.57

0.070.05

0.530.09

Scratchat

50µm/s-1h

6.521.99

0.480.03

3.532.21

0.030.05

0.520.06

indent-1h4.55

2.370.48

0.05

6.213.84

0.380.05

8.043.54

0.560.05

BF33

Scratchat

10µm/s-1h

6.110.51

0.650.009

2.161.31

0.130.009

0.780.01

Scratchat

50µm/s-1h

5.710.53

0.710.006

1.641.38

0.150.02

0.860.02

Indent-1h6.13

1.560.74

0.05

Indent-1h7.12

1.560.77

0.05

Page 57: Tribological studies on the surface of glasses by lateral

4.1. Studies at low load scratching in plastic regime of deformation 43

are consistent with previous reports (Vickers indentation [108,153,156,157]).

The biggest part of recovery stems from the face-to-face distances of residual

indents, as seen in cross-pro�le data in Fig. 4.3.

Silica with the lowest packing density of ions has the highest potential

for compaction under quasi-isostatic load [103]. The pile-up volume which

is observed in the present series of indentation experiments is small in com-

parison to Vickers indentation experiments [108,156]. This is related to the

inherent di�erences in the tip geometry. Small amounts of pile-up indicate

that the major deformation mechanism when using the Berkovich tip is den-

si�cation. Also the volume recovery values for silica and BF33 are relatively

close. This is not surprising considering the fact that both glasses are anoma-

lous [112]. Despite this, the borosilicate glass has a somewhat higher packing

density than silica and a higher number of non-bridging oxygen species, thus,

a slightly lower recovery ratio. SLG on the other hand has the lowest recov-

ery ratio due to its more dense structure, resulting from the presence of a

signi�cant amount of network modifying ions.

AFM top-view images of scratch grooves are presented in Fig. 4.4. Again,

the observed recovery ratio is the highest for silica and the lowest for SLG.

The pile-up volume is considerably smaller than the scratch groove volume,

indicating major densi�cation also in the scratching process. At this point,

we �nd no big di�erence between the overall recovery ratio of scratches and

indents. However, a signi�cantly higher amount of pile-up is observed fol-

lowing scratch deformation. This implies that di�erent deformation modes

are activated in the scratching process as compared to normal indentation,

further con�rmed by notable di�erences in the composition-dependence of

normal hardness and scratch hardness [150]. It is known that the shear de-

formation mode is more pronounced during lateral deformation [8], resulting

in somewhat lower volume recovery after annealing.

The annealing time and deformation-rate dependencies of VR for silica

are indicated in Fig. 4.5(a). The height of the pile-up region increased with

increasing treatment time and the volume recovery ratio almost doubles,

i.e., from approximately 0.4 to 0.8. Within the small variations of scratch

speed we considered here, this observation is independent of the rate of

deformation, Fig. 4.5(b).

Page 58: Tribological studies on the surface of glasses by lateral

44 Results and discussions

Figure 4.4: AFM top-views and cross-pro�le scans (positions indicated in

the AFM micrographs) of the residual scratch grooves after Berkovich edge-

forward scratching, before (left) and after (right) annealing at 0.95Tg, shown

for (a) silica, (b) BF33, and (c) SLG. Annealing times were 1h for BF33 and

SLG, and 2h for silica.

In order to verify the e�ect of stylus geometry on deformation recov-

ery, an additional experiment was conducted on BF33 in face-forward tip

con�guration (due to strong wear on the diamond tips in this set-up, the ex-

periment was not repeated for the other glasses). The resulting cross-pro�le

data are shown in Fig. 4.6. No considerable change of scratch groove volume

before heat treatment was observed for the two di�erent tip con�gurations

when using identical normal loads. However, after 1h of thermal anneal-

ing, the sink-in and pile-up volumes were slightly larger when the tip was in

Page 59: Tribological studies on the surface of glasses by lateral

4.1. Studies at low load scratching in plastic regime of deformation 45

Figure 4.5: (a) Cross-pro�le data of scratches on SiO2 before and after an-

nealing at 0.95Tg for 1 and 2h. (b) Recovery ratio versus annealing time for

di�erent scratch rates.

edge-forward orientation. This results from the reduction of contact surface

when using the sharp apex of tip for scratching as compared to face-forward

orientation. For similar scratch depth, the edge-forward con�guration pro-

duces stronger plastic �ow (analogous to normal indentation with a sharper

indenter). As a consequence, the recovery ratio VR- is smaller than for face-

forward testing. Di�erences in VR+ are a likely result of variations on the

stress distribution, producing a larger compressive zone at the scratch front.

In Fig. 4.7, we compare the volume recovery data of lateral and normal

indentation deformation. On �rst view, we �nd that scratch grooves and

indents have very similar overall recovery values, dependent on glass type. In

order to avoid chipping and microabrasion [158], the present scratch grooves

were generated at much lower normal load (17-30 mN) than the indents

(330-400 mN). However, the scratches experienced additional lateral load

components, approximately ∼ 1/4 of FN for a typical friction coe�cient of

0.25. For weaker loading, there is stronger recovery, in particular VR-, in

very good accordance with normal indentation studies done by Yoshida et

al. [159].

The value of VR- is decreasing with increasing normal load for SLG.

Overall, the highest degree of densi�cation and recovery is found for the

anomalous glasses (silica and BF33). The recovery ratio of the pile-up re-

Page 60: Tribological studies on the surface of glasses by lateral

46 Results and discussions

Figure 4.6: Cross-pro�le scans over a scratch groove on BF33 before and

after 1h annealing at 0.95Tg for face-forward and edge-forward orientation

of the Berkovich tip.

gion VR+ is relatively small for all three types of glasses. As expected, the

experiments carried-out at lower loads resulted in stronger pile-up recovery.

The observed trends in volume recovery and, hence, densi�cation during

scratch deformation exhibit a pronounced correlation with Poisson's ratio.

Also here, the observation is very similar to normal indentation [108]. The

densi�cation contribution to total indentation deformation decreases when

increasing Poisson's ratio from 0.153 for silica to 0.192 for BF33 and further

to 0.234 for SLG. This re�ects the relation between Poisson's ratio and the

atomic packing density [103]. In the anomalous glasses (silica and BF33),

the main mechanism of plastic deformation is densi�cation. Thus, a larger

fraction of the scratch-induced deformation is recovered upon heat treatment.

In the normal glasses (SLG), shear �ow plays a considerable role, re�ecting

in a lower recovery ratio. The network modi�ers may be understood as

providing �easy-slip �paths through a rigid covalent network [43, 160, 161],

leading to enhanced pile-up. Consequently, SLG exhibits pronounced pile-

up generation already during initial deformation, with little further recovery

(VR+). In BF33, pile-up extends during annealing, most probably as a

geometric result of stress relaxation.

Page 61: Tribological studies on the surface of glasses by lateral

4.1. Studies at low load scratching in plastic regime of deformation 47

Figure 4.7: Recovery ratios as a function of (a) loading values and (b) Pois-

son's ratio for all types of glasses. The dark symbols show the values ob-

tained in this study and the red symbols represent the values in the study

by Yoshida et al. [159]. (For interpretation of the references to colour in this

�gure legend, the reader is referred to the web version of this article).

4.1.2 Rippling inside the scratch groove of vitreous silica

During experiments of previous studies, it was noticed that some ripples were

formed inside scratch groove and within the plastic regime of scratching. One

scratch at loading rate of 10 µm representing such phenomenon is shown in

Fig. 4.8.

Since the formation of wavy patterns accompanying the motion of a lo-

calized object sliding on a compliant surface is a basic but not clearly under-

stood phenomenon occurring on very di�erent length scales, studying such

phenomena inside the scratches was motivation of this study. Examples of

this phenomena are the moguls formed in alpine skiing [162], the repeated

impressions left in standard scratch tests on polymers and glasses [163,164],

and the ripples observed when polymer [139], ionic crystal [165], and semi-

Page 62: Tribological studies on the surface of glasses by lateral

48 Results and discussions

Figure 4.8: An atomic force microscopy image of a part of one scratch per-

formed at scratching rate of 10 µm and increasing normal load of 10 mN

showing ripples in the scratch groove.

conductor [166] surfaces are scraped with ultralow normal forces of few tens

of nN.

In this section, an unambiguous evidence and quantitative data on the

occurrence of such regular surface rippling on silica under concentrated loads

of few mN is provided. From the velocity dependence of the ripple repetition

rate, this phenomenon is consistently associated to the stick-slip motion of

the indenter caused by periodic failure of the glass surface and deceleration

of the plowing motion.

The rippling phenomena can be accurately investigated in a variety of

controlled conditions (e.g., loading, scan velocity, and temperature) using

atomic force microscopy (AFM) only [137,138]. As an example, it has been

recently demonstrated that the formation of nanoripples on polymers can be

understood as a combined e�ect of the time-increasing shear stress imposed

by the slider (i.e., the sharp probing tip) and the viscoplastic response of

the material [29]. In this way, the rippling process appears to be ruled by

�ve parameters: normal force FN, indentation rate dz/dt, scan velocity v, tip

geometry, and lateral contact sti�ness k.

If the indentation pit has a Gaussian cross section of half width σ, sur-

rounded by a circular rim of comparable size, stick-slip motion accompanied

by ripples is expected if the inequality FN(dz/dt) > kσv holds, otherwise the

tip will move on continuously. Since the contact area between two solid sur-

faces sliding past each other usually consists of a multitude of tip junctions

which are continuously formed and broken at distinct sites, gaining knowl-

Page 63: Tribological studies on the surface of glasses by lateral

4.1. Studies at low load scratching in plastic regime of deformation 49

edge on regular wear patterns generated by a single tip, not necessarily on

the nanoscale, is crucial for modelling and controlling the response of moving

parts of mechanical components, as well as for fundamental aspects.

In this study, a combination of AFM and nanoindentation techniques is

adopted to investigate the results of stick-slip motion in the early stages of

mechanical contact on the surfaces of inorganic glasses with characteristic

depth and length scales in sub-µm range. Under these conditions, surface

deformation is at the origin of material strength and mechanical failure,

whereby initial �aws act as stress ampli�ers which reduce the applicable

strength by several orders of magnitude [105].

It should be noticed that, while surface scratch tests are routinely per-

formed on glass surfaces, the vast majority of such studies provide only qual-

itative and empirical information on scratch regimes where signi�cant crack-

ing or abrasion occur [63,105], although formation of patterns with repetition

distances of hundreds of µm was also reported in earlier investigations in the

high-loading regime [167]. The elastic-plastic regime, on the other hand, has

been addressed only recently, following major improvements in nanoinden-

tation technology, from the perspective of grinding performance [141, 142],

or scratching hardness [98, 150, 168]. In this regime, the fundamental rela-

tion between stick-slip and topography of the scratch grooves has not been

considered thoroughly so far.

An AFM image of a typical wear scar on the silica glass surface is shown

in Fig. 4.9(a). The scar has a width of about 3.2 µm and a depth of 250

nm. On the upper edge of the scar a modest pile-up of material is observed

with width and height of about 300 nm and 20 nm, respectively.

A cross section across the scratch groove [Fig. 4.9(b)] shows that the

transverse pro�le is not perfectly V shaped. The slope increases continuously

from 7°to 14°when approaching the axis of the groove. Only the limit value is

consistent with the geometry of the Berkovich indenter, which is expected to

cut the specimen with an angle α of about 15°with respect to the horizontal

plane. As shown in appendix (Fig. F1) both width and depth of the scars

increase with the normal force FN, whereas these quantities are not found

to change signi�cantly with the scan velocity v.

Page 64: Tribological studies on the surface of glasses by lateral

50 Results and discussions

Figure 4.9: (a) AFM topography of a silica glass surface previously scratched

(left to right) with a normal force of 30 mN and a scan velocity of 10 µm/s.

Set point: FN = 1.8 mN. (b) Cross section along the light blue line in (a).

(c) Simulated herringbone pattern obtained from the simple repetition of

Berkovich geometry every 350 nm without relaxation e�ects.

The accompanying AFM error signal maps [one of which is shown in Fig.

4.10(a)] clearly reveal the existence of a herringbone pattern across the whole

length and width of the scars.

The constituent ripples are tilted by 62°with respect to the axis of travel,

consistently with the triangular geometry of the Berkovich tip (projected

wedge angle of 120°). Several polishing lines, apparently unrelated to the

herringbone structure, can also be seen. The repetition distance (period)

λ of the ripples is not uniform. The average value of λ can be estimated

from 2D self-correlation of the error signal [Figs. 4.10(b) and 4.10(d)] or,

alternatively, from the 1D fast Fourier transform (FFT) along each horizontal

scan line [Figs. 4.10(c) and 4.10(e)]. The average values of λ coincide, and

FFT allows us to conclude that these values (220 ± 6 nm in the case of Fig.

4.10) are the same on all longitudinal sections (except for the central region

where some impurities are present).

Page 65: Tribological studies on the surface of glasses by lateral

4.1. Studies at low load scratching in plastic regime of deformation 51

Figure 4.10: (a) AFM error signal across the wear scar in Fig. 4.9(a). The

green arrows highlight two polishing lines. Frame size: 10×6 µm2; (b) 2D

self-correlation and (c) 1D-FFT along longitudinal direction extracted from

the region corresponding to the herringbone pattern in (a); (d),(e) cross

sections along the blue lines in (b) and (c), respectively.

As shown in the appendix (Fig. F2) the ripples can be also recognized in

longitudinal sections of magni�ed topography images. Their period is con-

sistent with the estimation in Fig. 4.10. Additionally, their rms amplitude

is 0.81 nm, corresponding to a corrugation on the order of 2 nm. For refer-

ence, ripple occurrence was veri�ed independently by confocal laser scanning

microscopy (Zeiss LSM): see Fig. 4.11.

The scratch test was repeated for six values of v between 10 µm/s and

500 µm/s while keeping FN at 30 mN. Each scratch groove was imaged at

di�erent (nonoverlapping) locations and the average period λ was calculated

as described before. As shown in Fig. 4.12, λ is found to increase almost

linearly with v. From linear regression it is possible to estimate that λ =

λ0 + t0v, with λ0 = (207 ± 8) nm and t0 = (1.07 ± 0.15) ms. This means

Page 66: Tribological studies on the surface of glasses by lateral

52 Results and discussions

Figure 4.11: Laser scanning microscopy image of a wear groove obtained in

the same conditions of Fig. 4.9(a). Scale bar of top view: 20 µm

that the period increases consistently across the range of applied velocities.

The corresponding increase of the rms roughness is more irregular- it varies

within 0.6 nm and 2.2 nm.

Additional information was obtained from the time variation of the in-

dentation depth and lateral pro�les acquired in-situ while the nanoindenter

was scratching the glass surface. As shown in Fig. 4.13(a) the average depth

of the wear groove was z0 = 460 nm, well above the value measured by AFM.

Together with the observation that the cross section is not V shaped, we

can con�rm that the glass signi�cantly recovered during unloading, in line

with previous indentation measurements (without scratching) based on Vick-

ers tips [169]. This behavior, which is not observed in other common glasses

such as soda-lime silicates, can be attributed to the very open structure of

silica glass and the strength of Si-O-Si bonds, facilitating elastic recovery.

From the value of the average lateral force FL = (4.40 ± 0.05) mN during

scratching [Fig. 4.13(b)] a coe�cient of friction µ = 0.15 is estimated. The

value changes only slightly when the velocity is increased (z0 = 475 nm

and FL = 4.5 mN at the maximum value of v = 500 µm/s). Noteworthy,

the corresponding in-situ rms value of the indentation depth, 0.87 nm, is

only slightly larger than the rms value of the ripples pro�les after scratching

Page 67: Tribological studies on the surface of glasses by lateral

4.1. Studies at low load scratching in plastic regime of deformation 53

Figure 4.12: Velocity dependence of the ripples period as measured by AFM

ex situ after scratching with a normal force FN = 30 mN (blue dots) and

linear �t of the experimental data points (red curve).

(0.81 nm), and the rms value of the lateral force ∆FL = 0.06 mN recorded

on line. Regular variations of these quantities corresponding to the period

of the ripple pattern cannot be distinguished with the lateral resolution of

the available setup (100 nm).

Important for the discussion, the indentation depth increases asymptot-

ically when the normal load is applied [as estimated from the red curve in

Fig. 4.13(a)]. The penetration rate dz/dt is indeed found to change from 175

nm/s [corresponding to the black arrow in Fig. 4.13(a)] to 50 nm/s (blue

arrow), with a characteristic time of 4.3 s. When scanning is resumed (at

the point shown by the green arrow) the penetration rate abruptly increases

to 250 nm/s for a short time of about 0.2 s. This means that the pressure

exerted by the nanoindenter is not balanced by the modi�ed surface pro�le,

and the latter is still evolving when the scratching process begins. Addi-

tionally, the lateral force is found to increase continuously when the scan is

resumed, as shown by the curve in Fig. 4.13(b). From the slope of the initial

part of this curve an e�ective lateral sti�ness k = 2.8 kN/m is estimated.

The saturation value of FL = 4.40 is reached quickly, within a time frame of

0.2 s.

Page 68: Tribological studies on the surface of glasses by lateral

54 Results and discussions

Figure 4.13: (a) Time variation of the indentation depth while scratching

with v = 10 µm/s. Scanning was stopped at the time t = 46 s (when the

normal load of 30 mN was applied, black arrow) and resumed at t = 51 s

(blue arrow). A steady state is reached slightly after (green arrow). The red

curve is an exponential �t of the data points in this time frame. (b) Lateral

force vs. indenter position when the scratching process initiates (at x = 20

µm). The positions indicated by the blue and green arrows correspond to

the times in (a).

With the information provided by the combined nanoindenter and AFM

measurements, it is possible to build up a reasonable interpretation for the

surface rippling phenomenon so observed. The time spent by the Berkovich

tip in each dip of the herringbone pattern is easily estimated as λ/v = 22.0

ms. During this time, the tip penetrates the surface by an additional depth

△z = (dz/dt)(λ/v) = 5.5 nm at most. This is larger than the corrugation

of the ripple pattern, as measured by AFM, which again suggests elastic re-

covery (with possible relaxation e�ects). Assuming that the tip apex sticks

to an indentation site while the tip is pulled along the x direction with in-

creasing lateral force FL, a point is eventually reached at which the glass

is not able to resist this force and the tip suddenly slips laterally. With a

cross-sectional area A of the wear groove of 0.74 µm2 [as estimated from the

scan width in the AFM images and the indentation depth in Fig. 4.13(a)]

and the lateral force value of FL = 4.4 mN measured in Fig. 4.13(b), the

corresponding hardness FL/A = 5.9 GPa, in line with recent observations

of the lateral hardness of silica [150]. The average energy dissipated in a

slip event can be also estimated as Ediss = FLλ = 0.97 nJ. This value cor-

Page 69: Tribological studies on the surface of glasses by lateral

4.1. Studies at low load scratching in plastic regime of deformation 55

responds to breaking of only 6% of chemical bonds in the removed volume

Vslip = Aλ = 0.163 µm3 (assuming a density ρ = 2650 kg/µ3 and an av-

erage bond energy of 621.7 kJ/mol [170]), suggesting that the ripples are

associated to nanofracture processes and not to glass �uidization caused by

local heating (which has been reported on larger scales in the context of laser

chemical vapor deposition [171]). Interestingly, for regular scratching within

the elastic-plastic domain, it has been also reported that the work of defor-

mation which is required for the creation of the permanent scratch groove

reaches about one tenth of the corresponding volume energy of silica [150].

To answer the question on what determines the distance travelled by the

tip before it stops again, i.e., the ripple period λ, two di�erent assumptions,

at least, can be made. At increasing velocities, the tip is expected to indent

less deeply the glass surface, reducing the lateral force required to induce the

slip. However, one would expect lower values of λ in this case, as observed on

polymer surface ripples slightly scraped by AFM [29], but in clear contrast

with the results in Fig. 4.12. This di�erence is due to the fact that the

indenter penetrates the surface much more deeply in the present case, and,

as proven by Fig. 4.13(a), it does not retract much more than 2 nm in the slip

phase. The duration tslip of the latter may also not be negligible, as compared

to the stick phase duration, tstick. As proven below, this leads to a linear

increase of λ with v. To this end let us consider the mechanical response

of a rigid object with mass m pulled by a spring with sti�ness k along a

surface. The sliding is slowed down by a constant kinetic friction force Fk

(which, in contrast to more complex numeric models of velocity-controlled

stick-slip [172], is simply assumed to be constant) and it becomes possible

only if the spring force overcomes a threshold value Fs > Fk, corresponding

to the static friction force. Under these general assumptions, it can be proven

(see appendix) that the two phases last for a time

tslip =2

ω0arctan(

vcv), and tstick =

2vcω0v

(4.2)

Page 70: Tribological studies on the surface of glasses by lateral

56 Results and discussions

Respectively, where ω0 =√︁k/m is the resonance frequency of the system

and υc = ω0(Fs-Fk). If v ≪ vc then tslip ≈ π/ω0 and the repetition distance

of stick-slip motion is λ = v (tstick + tslip)= λ0 + t0v, where

λ0 =2(Fs − Fk)

k, and t0 =

π

ω0(4.3)

This linear response (with initial o�set) perfectly matches the experi-

mental results in Fig. 4.12, after noticing that, in steady conditions, λ is

the same as the ripple period. With the values of λ0 and t0 from Fig. 4.12,

and k from Fig. 4.13(b), it is also possible to estimate values of 0.7g for

the e�ective mass m of the indenter and 0.28 mN for the di�erence ∆F =

Fs-Fk. The orders of magnitude of both values are consistent with the mass

of the stylus terminated by the Berkovich tip and the rms value of FL while

scratching (0.05 mN, comparable to the peak value for a perfect triangular

wave). While it is di�cult to draw more conclusions on the e�ective mass m

(depending on the nanoindenter design), the discrepancy in the values of ∆F

may indicate a smoothing of the transition between stick and slip phases,

reducing the lateral force (see Ref. [165] for the observation of this e�ect in

atomic-scale stick-slip measurements). This is also suggested by the blunt

features in the herringbone pattern, which, in Fig. 4.9(c), are compared with

the sharper pattern that would result from the simple geometric repetition

of the Berkovich pro�le every 350 nm. In this context it is also interesting to

observe that the very end of the scratch, when seen from above, also appears

rounded, as showed in the appendix (Fig. F4). This is a further con�rmation

of partial surface recovery after scratching.

4.2 Studies at ramp load scratching in plastic regime

of deformation

The focus of the studies in section 4.2 has been to look more deeply into dif-

ferent aspects of deformation within the constant low load scratching regime

of silicate glasses. From role of densi�cation and shear �ow in lateral inden-

tations to rippling phenomenon within the scratch groove were discussed.

However, it was seen in the Fig. 4.1 that in higher scratching loads other

regimes of deformation take place. Hence, ramp load scratching can provide

worthy information not only in higher loads of scratching, but also during

Page 71: Tribological studies on the surface of glasses by lateral

4.2. Studies at ramp load scratching in plastic regime of deformation 57

transition from elastic to plastic regime on the surface. Such experiments

have been mostly limited to soda-lime silicate glasses [1, 7, 8]. The next

two studies report deformation behaviour of vitreous silica in a quantita-

tive aspect and one metallic glass when applying higher loads in ramp load

scratching experiments.

4.2.1 Statistical analysis of microabrasion onset in vitreous

silica

It is well known that the structural reactions which underlie damage in�iction

are signi�cantly complex [151]. Di�erent regimes occurring during scratching

experiments have been studied in the last years [1,7]. However, a quantitative

study which shows the e�ect of di�erent parameters such as load, scratching

velocity, indenter geometry and so on on the deformation behaviour of glasses

was mostly missing and hence this study was designed.

This investigation focuses on lateral force analyses during scratching of

vitreous silica in an e�ort to obtain increasingly quantitative information

on the scratch resistance of vitreous silica. For this, the e�ect of scratching

velocity (loading rate) on the onset of micro-cracking and chipping is stud-

ied. The correlation of in-situ recordings of friction forces with post mortem

imaging of the scratch enables the identi�cation of onset points for scratch-

induced fracture events and microabrasion. This is subsequently evaluated

through Weibull statistics.

Data analysis

The apparent coe�cient of friction, µ, is approximated from the ratio of

lateral and normal load,

µ =LL

LN(4.4)

This assumes a fully plastic contact in which the progressing indenter is

opposed by the resistance of the material to be removed in its wake (Fig.

4.14).

As a quantitative measure of scratch resistance, the occurrence of the �rst

instantaneous cracking event during scratching was analysed. This analysis

Page 72: Tribological studies on the surface of glasses by lateral

58 Results and discussions

Figure 4.14: Evolution of lateral force during scratching (lateral displace-

ment) while increasing the normal load from 0.05 mN to 300 mN at a scratch-

ing velocity of 500 µm/s. The �rst 200 µm of lateral displacement correspond

to the pre-scan area (in which a normal load of 50 µN is applied). The crack-

ing region is highlighted: arrows mark individual cracking (chipping) events;

the onset of microabrasion is marked with a dashed line.

was performed on recordings of lateral force and apparent friction coe�cient

versus lateral displacement and indentation depth, respectively. In a typical

such scan (Fig. 4.14 or 4.15(a)), an approximately steady or, in the case of

LL analysis, even linear increase is observed in both parameters with increas-

ing normal load during scratching. At a certain stage, this steady evolution

is interrupted by sharp pop-ins. These are assigned to fracture events, lead-

ing to sudden bursts at the progressing scratch tip. As illustrated in Fig.

4.14, the occurrence of these bursts correlates with the post mortem scratch

pattern, where the onset of microabrasion corresponds to the onset of strong

discontinuity in the plot of friction coe�cient versus lateral displacement.

The corresponding value of LL was then obtained by extrapolation of the

initial regime of the plot (plastic regime, Fig. 4.15(a)) across those pop-ins

and taken as the lateral onset load for microabrasion. Besides the onset of

microabrasion, further features can be detected in the data shown in Fig.

4.15(a-b). For one, there is the onset of chipping where there appears to

be a deviation between the in-situ observation of friction (or lateral load)

and the post mortem consideration of the scratch pattern. This indicates

Page 73: Tribological studies on the surface of glasses by lateral

4.2. Studies at ramp load scratching in plastic regime of deformation 59

Figure 4.15: (a) Scratch pattern for fused silica at a scratching rate of 50

µm/s and a normal load which increases from zero to 300 mN, using an

irregular diamond edge for scratching. (b) is a representation of the corre-

sponding variation in the apparent friction coe�cient.

that similar to normal indentation, the initial radial (chevron) cracks appear

during unloading, i.e., after the scratching tip has passed the speci�c point

of occurrence. Such cracking events can therefore not be detected in-situ,

although they may cause weaker artifact features on the subsequent plot of

LL. The concrete veri�cation of the onset point of microabrasion was done

based on the appearance of pop-ins in the in-situ chart as in the example of

Fig. 4.14, and with the help of microscopic images as in Fig. 4.15(a). In all

three methods, an error bar of 5± µm was applied on the obtained value of

characteristic displacement.

Data on the onset of microabrasion were analysed in the form of a Weibull

distribution so as to obtain statistical information on the scratch-induced

cracking behavior of the material. The Weibull distribution is the most

widely used distribution to explain the distribution of �aws in brittle ma-

terials. This distribution not only provides a simple graphical solution, but

it is also useful when there are inadequencies in the data. For example, the

technique works with small samples and it is possible to identify the mixtures

of failures, classes and modes, which was one of the subjects of this study.

Page 74: Tribological studies on the surface of glasses by lateral

60 Results and discussions

The Weibull analysis is based on the weakest link theory according to

which the fracture behavior of a material is linked to the most signi�cant

defect within the material. That means that the �aw that most concentrates

the applied stress causes failure. To make this connection, it is necessary to

assume that fracture results from propagation of a �aw that is an extreme

in a distribution of �aws. These �aws have some distribution function in

relation to stress. The cumulative Weibull probability function is

Pf = 1− exp

[︃−(

σ − σuσθ

)

]︃m(4.5)

where Pf is the probability of failure at or below a given stress σ, σu is a

threshold parameter which represents the minimum stress below which a test

specimen will not break, σθ is the scaling parameter, taken as the character-

istic strength and dependent on specimen size and experiment con�guration,

and m is the Weibull modulus and determines which number of the family

of Weibull failure distributions best �ts or describes the data. Setting σu to

zero and taking the double logarithm of the resulting two-parameter Weibull

distribution yields

ln

[︃ln

(︃1

1− Pf

)︃]︃= m lnσ −m lnσθ (4.6)

The probability value of Pf is obtained through Benard's median rank

approximation,

Pf =i− 0.3

n+ 4(4.7)

where i is the rank of each data point in order of ascending LL and n is the

total number of scratches per experiment. Median ranks are used to obtain

an estimate of unreliability of each failure. It is the value that probability

of failure Pf should have after ith failure out of a sample of n components,

at a 50% con�dence level. 50% is the best estimate for unreliability and this

means that half the time the true value will be greater than 50% con�dence

estimate and in the other half the true value will be smaller than the estimate

[56,173�175]. After linearizing Eq. 4.6, m is obtained from the slope and σθ

from the intercept.

Page 75: Tribological studies on the surface of glasses by lateral

4.2. Studies at ramp load scratching in plastic regime of deformation 61

Phenomenology

In the following, the onset of microabrasion was studied by optical analysis

and the general phenomenon of sequential plastic deformation, chipping and

micro-abrasion were con�rmed, as in Fig. 4.1. In the given example in Fig.

4.15, the onset of microabrasion is clearly visible at a scratch length of ∼450µm. In the plot of µ versus displacement (Fig. 4.15(b)), this corresponds

to the onset of frequent pop-ins without recovering continuous friction. The

overlap between in-situ variations in µ and post mortem optical inspection,

for the microabrasive regime, has two consequences. For one, it indicates

suitability of the observation of µ for accurately quantifying the onset of

microabrasion. Secondly, it indicates that the pop-ins which are observed in

the microabrasive regime correspond to practically instantaneous cracking.

For broader veri�cation, the characteristic onsets of microabrasion (OM)

of all experiments from the present study such as obtained from µ and from

optical inspection, respectively, are plotted against each other in Fig. 4.16.

As an example, there is good accordance between the two ways of observ-

ing OM, i.e., a linear correlation with slope ∼ 1.014 and Pearson product-

moment correlation R ∼ 0.8717 for the chart of Fig. 4.16a (the low value of

R is a result of two individual extreme outliers).

More detailed inspection of Fig. 4.15(b) reveals the occurrence of indi-

vidual pop-ins already before microabrasion, i.e., in the region of chipping.

Here, the points of occurrence do not clearly correspond to post mortem

optical observations. Assuming that these individual pop-ins are not experi-

mental artifacts, this means that the underlying cracks are either too small to

be visually resolved, or that they do not occur on the surface instantaneously,

e.g., that they form sub-surface and/or during unloading in the wake of the

scratching indenter, so that the crack intersection with the scratch groove

which as identi�ed post morten does not correspond to the point of in-situ

observation of the disturbance of the moving indenter. The latter interpreta-

tion is supported by the corresponding observation that certain cracks which

are visible post mortem do not have a parallel pop-in event in the in-situ

scans. In the correlation graphs of Fig. 4.16, such events are the primary

reason for the extreme outliers. Direct inspection of in-situ recordings of

lateral force LL provides a very similar picture (Fig. 4.14). Here as well, in-

Page 76: Tribological studies on the surface of glasses by lateral

62 Results and discussions

Figure 4.16: Determination of the onset of microabrasion (scratch length in

µm) OM through di�erent methods: (a) Post mortem optical microscopy

and in situ observation of the apparent coe�cient of friction, and (b) optical

microscopy and in situ observation of the lateral force. In (c) the determi-

nation of lateral force is considered, i.e., as read directly during in situ scans

and as determined from the length at which OM was observed through the

apparent coe�cient of friction, µ, according to (a). The lines represent linear

correlation �ts with indicated slope and Pearson product-moment correlation

R. Data derive from 20 individual experiments for each of the 6 scratching

velocities. In (c), the three data points marked with an asterix were ex-

cluded from the linear �t. All lines represent linear �ts of the data, with

�tting parameters given in the respective panel.

dividual pop-ins are detected across the phenomenological regions of radial

cracking and chipping (Fig. 4.1).

The occurrence of the �rst pop-in correlates roughly with the transition

from the plastic regime to the regime of radial cracking at a displacement

of ∼ 400 µm (including the pre-scan of 200 µm, Fig. 4.15(a-b) and Fig.

4.14). The non-linear onset of µ in the �rst 200 µm of the scratch is caused

by non-linear LL. Interestingly, the onset of microabrasion corresponds to

the onset of linearity in LL and, consequently, in µ. It remains a question

of further study as to how this non-linearity correlates to, e.g., indentation

size e�ects and the strain-rate dependence of indentation response. Since in

the present experiment, a linear increase of LN is imposed on the system,

information which is provided through in-situ determination of µ or LL is

physically equivalent. Thus, determination of the onset of microabrasion

from scans of µ or LL should provide equivalent results. Individual pop-

Page 77: Tribological studies on the surface of glasses by lateral

4.2. Studies at ramp load scratching in plastic regime of deformation 63

ins are more clearly visible in the LL scan, especially for lower normal load

where they are not smeared-out through mathematical division. Vice versa,

the onset of microabrasion is better seen in the µ scan, which also leads to

better agreement with optical inspection, Fig. 4.16(a-b).

Over a series of experiments with increasing normal load, the scratching

distance and onset load of microabrasion are not constant. As an exam-

ple, such data are provided in Table 4.2 for a scratching speed of 50 µm/s

(corresponding to a normal loading rate of 15 mN/s).

This indicates a strong contribution of the material surface condition

[176] and/or experimental parameters (such as the proper orientation of the

tip in EF con�guration, Fig. 4.14) to the occurrence of scratch-induced sur-

face cracks, similar as with strength testing or determination of the load of

crack initiation through normal micro-indentation [104, 105, 177]. In some

cases, the responsible surface �aws can readily be detected by optical inspec-

tion, e.g., Fig. 4.17.

Figure 4.17: Post mortem optical microscopic image of a scratch generated

at a scratching speed of 10 µm/s under increasing normal load (3 mN/s).

The onset of microabrasion (marked) was observed at a normal load of 153.3

mN and a lateral load of 42.4 mN. In this example, the initiating surface

�aw was a scratch, probably induced during polishing (marked).

Following the above observations, statistical analyses were performed in

order to extract quantitative data on the scratching behavior of vitreous

silica. This included analyses of the probability Pf for the occurrence of

microabrasion. In Fig. 4.18 data are provided for LL at the onset of mi-

Page 78: Tribological studies on the surface of glasses by lateral

64 Results and discussions

Table

4.2:Onset

ofmicroabrasion

(OM)for

aseries

of20

experim

ents,scratching

vitreoussilica

atarate

of50

µm/s

with

normalload

increasingfrom

0.05mNat

arate

of15

mN/s.

Experim

entalerrors

are±5µm

onall

displacement

data,and

±0.01

mNon

allloads.

Displacem

entof

OM

(µm)

experim

entno.

Inspection

ofµ

Optical

inspection

Inspection

ofLL

LNat

OM

(mN)

LLat

OM

(mN)

1193

206273

81.922.1

2275

280266

79.519.9

3158

140190

56.915.7

4643

660499

15054.3

5439

370429

12832.3

6159

105130

38.910.8

7731

746131

39.058.6

8554

334513

15439.2

9111

180128

38.39.5

10779

720703

21155.4

11629

590601

18047.6

12555

593560

16842.7

13430

440431

12932.6

14208

210214

64.115.9

15631

600617

18549.5

16449

500486

14538.2

17550

580540

16243.3

18291

200209

62.616.7

19530

580513

15439.1

20605

525597

17947.1

average446

427.9401.5

162.134.5

Page 79: Tribological studies on the surface of glasses by lateral

4.2. Studies at ramp load scratching in plastic regime of deformation 65

Figure 4.18: Statistical analysis of the onset of microabrasion (OM) in vitre-

ous silica during lateral indentation. Depicted data present the probability

of the occurrence of microabrasion as a function of acting lateral load LL.

They are plotted for varying scratching velocity with gradually increasing

normal load (0.0 mN to 300 mN) as Weibull distribution with the proba-

bility term Pf according to Benard's median rank approximation. OM was

determined from in-situ lateral force measurements. Lines represent linear

�ts of the intermediate failure regime (�t data in Table 4.3). As a guide to

the eye, slopes of 1, 2 and 6 are also indicated.

croabrasion (including the data given in Table 4.2 for the case of scratching

at 50 µm/s, 15 mN/s).

Clearly, the probability of failure is higher at higher lateral loads. A min-

imum value of LL is required to start any abrasion, consistent with obser-

vations [60] on lateral cracking during unloading in quasi-static indentation

experiments. Data on Weibull modulus are summarized in Table 4.3.

Roughly, there are three regimes of failure (failure modes): the �rst fail-

ure mode occurs at relatively low load with relatively high slope (best seen

in the plots for scratching velocities of 10 µm/s and 100 µm/s, Fig. 4.18).

This mode can be attributed to the occurrence of major surface �aws and/or

Page 80: Tribological studies on the surface of glasses by lateral

66 Results and discussions

Table 4.3: Weibull parameters for failure modes I and II and varying speed

of scratching.Velocity (µm/s) m R-value for m

10 1.61 0.973

50 2.73 0.982

100 1.72 0.996

150 2.04 0.974

300 2.82 0.983

500 4.40 0.985

experimental perturbations. In particular, individual outliers at lowest load

indicate the occasional presence of a distinct, single disturbance or �aw.

They are thus not taken into account in the following quantitative evalua-

tion. A second regime is seen at very high load, visible only for high scratch-

ing rates (300-500 µm/s, Fig. 4.18). In this regime, the probability func-

tion levels-o� to an exponential equation, indicating constant rate of failure.

Such independence of mechanical failure on load (random OM) can be in-

terpreted as resulting from high overall surface quality (thus, generally high

abrasion resistance) with very occasional (laterally randomly distributed)

failure-inducing �aws. This means that a high scratching rate smears-out

the occurrence of OM in some speci�c samples with low surface defect den-

sity. Vice versa, it can also be concluded that some �aws are activated only

at low scratching rate: at high enough rate, the indenter is simply passing-by

some types of defects.

For further evaluation, at the moment, only the intermediate regime was

considered as marked in Fig. 4.18. For this regime, the Weibull modulus is

found in the range of roughly 1.6 - 4.4, increasing with increasing scratch

velocity and/or increasing loading rate. Hence, the underlying probability

of failure exhibits a compressed exponential or even Gaussian distribution

which is further compressed with increasing loading rate. Especially at low

scratch velocity, the values are somewhat below but still in the range of those

which are typically found in macroscopic testing of similarly prepared glass

samples, e.g., by ring-on-ring cracking [178]. On the one hand, this signi�es

the much lower tested volume. On the other hand, it also indicates that at

high enough scratching rate (and, thus, tested length), scratch-induced mi-

Page 81: Tribological studies on the surface of glasses by lateral

4.2. Studies at ramp load scratching in plastic regime of deformation 67

crocracking is similarly a�ected by the presence of surface �aws as is macro-

scopic cracking (notwithstanding the above arguments regarding the regime

of high load/high rate).

As noted above, with the exception of the experiment which was con-

ducted at 50 µm/s, the obtained Weibull moduli depend roughly linearly

on scratching velocity. Looking at the exact data (Fig. 4.18), the increase

in the value of m originates from an overall compression of the data. That

is, the underlying probability function is compressed on the low-load side,

leading to postponed activation of certain �aws at higher scratching velocity

and/or accumulation in individual cracking events rather than continuous

microabrasion. Data then catch-up at higher load, leading to a steeper pro-

�le in the Weibull plot. Vice versa, at lower velocity, more time is left for

defect activation and growth. Aside of individual outliers (in the low-load

failure regime, see above), the onset of microabrasion follows a normal or

even compressed exponential distribution. This re�ects a situation where

the proceeding scratch is intersecting randomly distributed surface defects

with a decreasing stress-dependence of their activation to form cracks. A sim-

ilar conclusion was drawn by for the case of soda lime silicate glasses [50,73].

Also for these, it was found that changing the scratching velocity a�ects the

cracking behavior. This was attributed to extended time interval over which

any one surface defect stays in a stressed state at lower scratching velocity,

so that its probability of growing into a visible surface crack increases. It

remains to be examined in future studies how this is related to subcritical

defect growth.

4.2.2 Scratching of metallic glass by a blunt indenter

So far, all the studies of this dissertation were concentrated on silicate glasses.

However, many industrial applications require materials with remarkable and

sometimes contradictory properties. For example, in the �elds of biomate-

rials (dental implants), micromechanics (gears) or in the �eld of jewellery

or watches (luxury watches), materials are needed that are hard, wear resis-

tant, bio-compatible, possess a high yield strength, while being deformable.

Such ideal materials do not exist yet. Metallic glasses have shown a high

potential for industrial applications due to their high strength, high elastic

limits, excellent corrosion resistance, and thermoplastic formability. On the

Page 82: Tribological studies on the surface of glasses by lateral

68 Results and discussions

other words, the combination of their structural and functional properties

make them potential candidates for applications where the use of conven-

tional materials has reached a limit of e�ectiveness. Following these reasons,

investigations on tribological properties of metallic glasses have been the

focus of many studies in the last decades [31�37].

Nonetheless, the dynamic behavior of nanoscratching in metallic glasses

which are relatively new materials in comparison with oxide glasses has not

been given much attention. Ramp load scratching seems an appropriate

method to study the tribological properties of metallic glasses and to col-

lect initial information about the surface resistance of these group of glassy

materials. Such information could likely be used to design constant load

scratching tests later which provide more precise information about the plas-

tic deformation mechanisms of di�erent groups of metallic glasses. Hence,

in this section, some aspects of deformation behavior of a metallic glass

(Zr55Cu30Al10Ni5) which has a good forming ability and a great potential

for industrial applications was studied by ramp load scratching.

These series of experiments were performed with a conical spherical in-

denter. The main goal of scratching experiments with a blunt indenter is to

outline a mechanistic framework for interpreting measurements performed

on elastic/plastic materials. Moreover, at low force levels, the stresses be-

neath a blunt indenter are below the elastic limit, and hence, the tribological

properties of material can be obtained in the absence of plasticity. Moreover,

at higher forces, responses in the transitional elastic/plastic and fully plastic

can also be probed. Selection of the spherical indenter is further motivated

by the recognition that the asperities that make contact during sliding of sur-

faces are more closely represented by protuberances with a constant �nite

curvature rather than ones with in�nitely sharp points [57,179�181].

Fig. 4.19 shows an scanning electron microscopic overview image of two

scratches performed at the maximum loads of 30 mN and scratching rate of

10 µm/s. The scratching direction is from right to left and along this route

the scratch grooves become deeper and wider by increasing the load. At

lower loads the scratch groove is so shallow that it creates a inconspicuous

impression.

Fig. 4.20 shows the load-displacement curve for indentation and scratch-

ing experiments performed at the maximum loads of 20-50 mN.

Page 83: Tribological studies on the surface of glasses by lateral

4.2. Studies at ramp load scratching in plastic regime of deformation 69

Figure 4.19: (a) An Scanning Electron Microscopy overview image of two

scratches performed at load of 30 mN and scratching rate of 10 µm/s. The

whole scratching length is not shown in the image. (b) A pro�le of displace-

ment into surface over the whole lateral length for a scratch at the same

conditions.

In P-h curves for indentations and prior to the �rst pop-in, the loading

curve exhibits a smooth and parabolic shape, which can be described by

Hertz's law for an elastic contact [182]:

P =4

3Er

√Rh3 (4.8)

Where Er is the combined elastic response of the indenter tip and the

glass specimen and h is the displacement into surface. If both contacting

Page 84: Tribological studies on the surface of glasses by lateral

70 Results and discussions

Figure 4.20: The normal load vs. penetratuin depth curves for indentation

(left) and scratching (right) experiments at normal load of 20, 30 and 40 mN.

All scratching experiments were performed at scratching rate of 10 µm/s.

bodies have a curvature, then R in the above equations is their relative radii

given by:

1

R=

1

R1+

1

R2(4.9)

By using a spherical indenter and during the indentation the load-displacement

curve often exhibits a burst of displacements that is called pop-ins and they

can well be seen in the Fig. 4.20. Serrated plastic �ow phenomenon has been

widely observed in BMGs under deformation-constrained loading modes such

as compression [183�192] and especially the �rst pop-in matter. Because

it shows the transition from elastic to plastic behaviour and are believed

Page 85: Tribological studies on the surface of glasses by lateral

4.2. Studies at ramp load scratching in plastic regime of deformation 71

to be the birth of shear bands in nanoindentation experiments on metallic

glasses [32, 189]. By this explanation, it can be noticed in Fig. 4.20 that

the applied load drops at the onset of pop-ins and the �rst pop-in in 4.20(b)

appear at the load of 12.1 mN. There was no pop-in for P-h curves of inden-

tation at the load of 10 mN which means the material showed a rather elastic

behavior. Though at 20 mN load for scratching experiment pop-ins are seen

on the P-h curves (Fig. 4.20(b)), the indentation experiments showed almost

no pop-ins until the load of 40 mN (Fig. 4.20(e)). The P-h curve obtained

at the load of 40 mN for indentation indicates that the deformation at this

load was mostly elastic too. However, around the maximum load, the �rst

pop-ins occurred (Fig. 4.20(e)). The charts of experiments at 50 mN are

not showed here due to the tip blunting and this matter will be discussed

further.

All the pop-in loads for indentation and scratching experiments were

estimated from the P-h charts in Fig. 4.20 and the results are shown in

Table 4.4. The dash lines represent the samples for which it was not possible

to read the pop-in loads clearly. The lateral loads corresponding to these

normal loads were also estimated and shown in the Table 4.4 and excluding

the scratch at 50 mN, they are between 0.13 and 1.6 mN, around one tenth

of the normal loads. The pop-ins appearance is load dependent which is in

agreement with other literatures [193�195]. Lower load activates a smaller

volume of material and such dependence of plastic deformation to sample

size has already been discussed in other literature [196]. Also the higher the

normal load, the higher goes the lateral load. The scattering of pop-in loads

at 50 mN could be related to tip blunting. After all, it is believed that if

the tip blunting would not have happened, we would have had lower pop-in

loads for 50 mN experiment due to higher plastic deformation, as described

before in section 4.1.1.

Fig. 4.21 shows such comparison more distinctively. Another interest-

ing observation is that even for indentations and scratching performed at

identical conditions the onset of plasticity is distributed over relatively large

range. This is in agreement with results of Refs. [32, 197].

For example, loads ranging from 7.84 up to 22.2 mN have been required

to observe the �rst pop-in, in the Cu48Zr48Al4 alloy, although all indentations

were performed under the same experimental conditions. This phenomenon

Page 86: Tribological studies on the surface of glasses by lateral

72 Results and discussions

Table 4.4: Normal loads and lateral loads at which the pop-ins appear for

each experimental load at 20, 30, 40, and 50 mN.

Maximum normal load of experiment (mN) Normal load at pop-in (mN) Lateral load at pop-in (mN)

Scratch at 20 mN 12.1 0.25

- -

- -

17.3 0.13

15.6 0.41

Indent at 30 mN 28.2

-

-

-

-

Scratch at 30 mN 11.9 0.3

12.5 0.19

13.2 0.16

12.5 0.63

- -

Indent at 40 mN 39.9

39.9

38.1

38.4

37.9

36.5

Scratch at 40 mN 19,4 1.6

16.6 1.1

- -

- -

- -

Indent at 50 mN 43.0

43.3

40.4

44.5

39.5

Scratch at 50 mN 33.9 2.44

24.5 2.35

30.0 3.43

23.5 2.55

- -

has at least partly been attributed to the spatial heterogeneity in the local

atomic con�gurations inherent to metallic glasses [197�203]. Yielding can

occur at any stress when the thermal energy is high enough, whereby the

probability for yielding increases exponentially with increasing load [204,

205].

Page 87: Tribological studies on the surface of glasses by lateral

4.2. Studies at ramp load scratching in plastic regime of deformation 73

Figure 4.21: Normal loads at �rst pop-ins appearance for indentations and

scratching (left) and lateral loads for scratching experiments (right).

Schuh and Lund [188] suggested that thermally assisted and stress-biased

yielding always exhibits a spread in yield strength. This is because the ther-

mal noise sometimes favors yielding and sometimes works against it. Yet,

in this case, more experimental data will determine the range more pre-

cisely. Furthermore, the �rst pop-ins in scratching experiments appear at

lower loads in comparison with indentation experiments. It can be con-

cluded that scratching is a more severe deformation mode than indentation

for the same normal load: the in�uence of the tangential load, and therefore

friction among others factors may contribute to the di�erences in the plastic

deformation mechanisms. On the other words, for the same loads, scratch-

ing is more likely able to make the material enter the fully plastic regime of

indentation. Moreover, the higher the load, the higher the number of bands

at the surface, to accommodate plastic deformation [206]. This phenomenon

has been discussed in section 4.2.1 through estimating the probability of

occurrence of a certain type of plastic deformation (in that case onset of mi-

croabrasion) by increasing the load and it was proved that such probability

increases by applying higher load on the sample surface.

It was mentioned that the load-displacement chart of scratching at 50

mN was not shown with other curves due to the tip blunting (the data were

not consistent and it was concluded that something may have occurred here).

The e�ect of normal load on coe�cient of friction (COF) was further studied

Page 88: Tribological studies on the surface of glasses by lateral

74 Results and discussions

to explain the reason. Fig. 4.22 shows the variations of COF along the

scratching distance and it is increasing slightly with the increase of normal

load along the scratch groove and drops in the end of scratch where the

indenter is slowly leaving the surface.

Figure 4.22: Coe�cients of frictions for scratches at loads of 20, 30, 40 and

50 mN. Each experiment was repeated 5 times.

The COFs are also increasing by increase of load from 20 mN to 50 mN.

The increase in Coe�cient of friction (COF) with the applied normal load

both at micro- and macro-scales have already been reported [193�195] and

it could be due to increasing amount of roughening or increase of plastic

deformation at higher loads [33]. The COF close to zero have been due to

the very weak contact between the indenter and surface as a result of low

loads. At the load of 40 mN, the last 3 experiments are showing higher

values than the �rst 2 and it is believed to be related to tip blunting at third

experiment. For the reason of tip blunting, the contact between indenter

and material surface has increased and this indicates an increase in friction

Page 89: Tribological studies on the surface of glasses by lateral

4.2. Studies at ramp load scratching in plastic regime of deformation 75

coe�cient in COF charts. For each load the COF values are closed to each

other for 5 experiments and are between 0 to 0.15 in between the friction

coe�cient estimated by applying load with spherical indenters on the surface

of glasses (fused silica µ < 0.065 at loads of up to 5 N [133], Borosilicate µ <

0.065 for loads between 0 to 6N [207]) and metals (Copper µ <0.25 for loads

up to 1N [4] and for Aluminm 0.4< µ <1 [57]).

In section 4.1.2 it was mentioned that some ripples had been observed

under the microscope that were further studied. In this study, some pat-

tern (semi-circular shape) was also observed inside the scratch groove. The

shapes start to appear in the groove at a certain load and continue to the end

of scratch, as it can be seen in 4.23(a). These patterns were further partly

investigated by Scanning Electron Microscope (SEM) and Atomic Force Mi-

croscope (AFM). Fig. 4.23 demonstrates SEM and AFM images and the

pro�le of patterns along the scratch groove at the load of 30 mN.

In Fig. 4.23(a), it is noticed that these semi-circular shapes have di�erent

frequencies along the scratch groove. Similar patterns have been observed

in some studies carried out on the glasses, however not deeply investigated

so far to authors knowledge [4, 134, 207, 208] and they were described to be

a result of indenter shape on the surface. But so far no study was found on

metallic glasses discussing such patterns. A good e�ort was done to estimate

the wavelengths of these patterns that seem to change with the loading (it

can be seen more obviously in Fig. 4.23). But due to the very shallow depth

of these structures (some tenth of nanometres in comparison with ripples in

section 4.1.2 which had depths in the dimension of hundreds of nanometres),

it was not possible to calculate the precise wavelengths (an example of these

calculations is presented in appendix 5). However, the scratching distance

at which these patterns start to appear was read from SEM images and the

load correlated to this point was determined directly from indenter data and

these were compared with pop-in loads which were estimated from charts of

Fig. 4.21 for the scratching loads of 20 and 30 mN (at the load of 20 mN

only one scratch showed the clear appearance of pattern in SEM images).

Fig. 4.24 shows this comparison.

Interestingly, the pop-in loads and the loads at which the patterns appear

in SEM images are not that far from each other and they are all between 12-

15.5 mN. Even for two di�erent scratching loads, the �rst pop-in is appearing

Page 90: Tribological studies on the surface of glasses by lateral

76 Results and discussions

Figure 4.23: a) An image of a scratch at the load of 30 mN obtained by

Scanning Electron Microscope b) and c) Atomic force microscopy images of

two marked sections of the same scratch in (a), d) and e) Center line pro�les

of AFM images in (c) and (d).

around the same load. From here, we may conclude cautiously that there

could be a relationship between appearance of pop-ins and semi-circular

shapes or in other words, the patterns occurrence is related to formation

of shear bands in material. However, more experiment in di�erent loads is

necessary to con�rm such phenomenon. Another observation is that the pop-

in read from the charts of indenter appear after patterns appearance from

SEM along the scratch length. There is roughly 20-50 µm distance between

pop-ins scratching distance read from indentation charts and SEM data.

The time between appearance of pop-in in indentation data and formation

of patterns on SEM images was estimated to be ∼ 3-5.5 s which is relatively

small considering the total scratching time of <72 s for the load of 20 mN and

<140 s for the load of 30 mN. To answer the question why is this happening

Page 91: Tribological studies on the surface of glasses by lateral

4.2. Studies at ramp load scratching in plastic regime of deformation 77

Figure 4.24: Comparison of pop-in loads read from indenter and the point

of appearance of patterns in SEM images for the scratching loads of 20 and

30 mN. An standard deviation of 5% µm should be taken into account for

the data.

and could it be related to estimation of distances in SEM images or it is

really a phenomena needs more investigations.

Page 92: Tribological studies on the surface of glasses by lateral
Page 93: Tribological studies on the surface of glasses by lateral

Chapter 5

Conclusions

The aim of this thesis was to examine the tribological properties of di�erent

glasses by applying lateral nanoindentation technique. A scratch pattern

produced by ramp load experiment on the surface of silica was the start

point of further studies. This pattern showed di�erent deformation regimes

taking place on the surface at low and high loads. While radial and me-

dian cracks appeared in low load regimes, lateral cracks and microabrasion

were taking place in higher loads. This study showed the potential of more

methodological investigations on various types of glasses. From here, two

study were designed in constant low load and two more in higher ramping

load scratching on more silicate glasses and one metallic glass.

Through relaxation experiments performed on the scratch-induced sur-

face of fused silica(anomalous), borosilicate and soda lime silica (normal)

glasses in low load regime, the role of compaction and shear deformation

was studied. It was demonstrated that structural compaction and shear

�ow occur also in the lateral deformation of a glass surface through scratch-

ing. Applying instrumented indentation with tangential displacement in the

elastic-plastic regime of scratch-deformation, a strong similarity was found

in the volume recovery ratio by post-annealing as compared to normal inden-

tation studies. Silica, borosilicate and soda lime silicate glasses follow the

same trend in terms of compaction recovery with Poisson's ratio as they do

in normal (quasi-isostatic) indentation. Therefore, the same distinction into

normal and anomalous glasses appears to be applicable. On the other hand,

inherent di�erences occur in the absolute presence of deformation modes

79

Page 94: Tribological studies on the surface of glasses by lateral

80 Conclusions

across all three types of glass. In particular, caused by shear deformation at

the apex of the employed Berkovich tip, pronounced material pile-up occurs

in scratching for normal loads which are about one order of magnitude below

reference experiments of normal indentation. This leads to an increase in the

e�ective friction coe�cient and a non-trivial correlation between the scratch

hardness and the normal hardness of glasses.

It was found that lateral indentations with moderate loading forces (up

to 30 mN) on the surface of fused silica form some rippling phenomenon.

This was further studied at di�erent scratching velocities (up to 50 µm3)and

was related to periodic stick-slip motion of the tip. The ripples period was

estimated and showed a rather linear dependency to the scratching velocity.

Since most sliding interfaces of technical interest are ultimately formed by

a multitude of contact points which are abruptly formed and broken, the

information so acquired can be of key interest for understanding much more

complex wear processes occurring on the macroscale. Speci�cally, one could

expect that not only the irregular time variation of the friction force in a

multiasperity sliding contact results from an ensemble of regular stick-slip

events, but the accompanying abrasive wear patterns are ultimately caused

by overlapping (and possible interference) of �ripple-like�features similar to

those reported here in the rather ideal case of an inverted pyramid vs �at

contact. Fourier analysis of the lateral force acquired while scanning may

also help to correlate the time evolution of the stick-slip mechanism to the

topographical features of the resulting surface patterns.

Ramp load scratching was recognized as an appropriate technique to

study the tribological properties of glasses. Since most studies so far focused

on obtaining qualitative information on the surface of glasses, instrumented

nanoindentation with ramp loading was employed for obtaining quantitative

information on the onset of scratch-induced microabrasion on silica glass. For

this, in situ evaluation of lateral force and friction coe�cient was compared

to post mortem optical inspection, following edge-forward scratching with a

Berkovich indenter. Statistical analysis indicated two underlying probability

functions for the occurrence of microabrasion, i.e., the probability for the

propagating scratch to hit a surface �aw and the probability that such an

event causes an observable micro-crack. Dominance of the former follows

an exponential function, re�ecting a purely random distribution with load

independent probability of failure. It was observed only at high scratching

Page 95: Tribological studies on the surface of glasses by lateral

Conclusions 81

velocity after passing a certain normal load. For the latter, the Weibull mod-

ulus was found to increase with increasing scratching velocity, i.e., from ∼ 1.6

to 4.4. Here, low Weibull modulus at low load was attributed to the increas-

ing time of local strain, which leads to a reduction of the load-dependence of

micro-cracking. For silica and employing the present experimental approach,

the critical lateral load for microabrasion (50th percentile) is around 30-40

mN. This value is very probably dependent on extrinsic parameters such as

ambient humidity.

Further, ramp load scratching technique was used to obtain tribological

information on the surface of (Zr55Cu30Al10Ni5) metallic glass which showed

a great potential for industrial applications. Here as well, it was con�rmed

that the scratching is a more severe deformation mode in comparison with

normal indentation. This was argumented through the study of �rst pop-in

P-h charts. The friction coe�cient estimated from indentation data was be-

tween 0 and 0.15, well between the friction coe�cients of glasses and metals.

Further scanning electron microscopy study of the scratches on the surface

showed that there might be a relationship between the start of shear defor-

mation and semi-circular pattern that were observed on the surface.

Page 96: Tribological studies on the surface of glasses by lateral
Page 97: Tribological studies on the surface of glasses by lateral

Zusammenfassung

Das Ziel dieser Arbeit war es, die tribologischen Eigenschaften von ver-

schiedenen Gläsern mit Hilfe der lateralen Nanoindentationstechnik zu unter-

suchen. Ein Kratzmuster Rampenbelastungsexperiment auf der Ober�äche

von Quarzglas erzeugt Ausgangspunkt für weitere Untersuchungen. Dieses

Muster zeigte verschiedene Deformationsregime die auf der Ober�äche bei

niedrigen und hohen Belastungen statt�nden. Während radiale und mediane

Risse bei niedrigen Belastungen auftraten, traten bei höheren Belastungen

laterale Risse und Mikroabrieb auf bei höherer Belastung auf. Diese Studie

zeigte das Potenzial für weitere methodischen Untersuchungen an verschiede-

nen Glastypen. Von hier aus wurden zwei Studie in konstanter niedriger Be-

lastung und zwei weitere in höherer rampenförmiger Lastkratzens an weit-

eren Silikatgläsern und einem metallischen Glas durchgeführt.

Durch Relaxationsexperimente, die an der kratzinduzierten Ober�äche

von Quarzglas (anomal), Borosilikat- und Kalknatrongläsern (normal) Gläsern

im Niedriglastbereich durchgeführt wurden, wurde die Rolle der Verdich-

tung und Scherverformung untersucht. Es wurde gezeigt, dass strukturelle

Verdichtung und Scherung �ieÿen auch bei der lateralen Verformung einer

Glasober�äche durch Kratzen auftreten. Die Anwendung einer instrumen-

tierten Vertiefung mit tangentialer Verschiebung im elastisch-plastischen Regime

der Kratzverformung, wurde eine starke Ähnlichkeit im Volumenrückgewin-

nungsverhältnis durch Nachglühen im Vergleich zu normalen Indentations

Studien. Silikat-, Borosilikat- und Kalknatronsilikatgläser folgen dem gle-

ichen Trend in Bezug auf die Verdichtungserholung mit der Poisson-Zahl, wie

sie es bei normaler (quasi-isostatischer) Eindringung. Daher scheint die gle-

iche Unterscheidung in normale und anomale Gläser anwendbar zu sein. An-

dererseits treten inhärente Di�erenzen im absoluten Vorhandensein von Ver-

formungsmodi bei allen drei Glastypen auf. Insbesondere, verursacht durch

Page 98: Tribological studies on the surface of glasses by lateral

Scherdeformation am der Spitze der eingesetzten Berkovich-Spitze kommt es

zu einer ausgeprägten Materialanhäufung beim Kratzen für normale Belas-

tungen, die etwa eine Gröÿenordnung unter Referenzversuchen der normalen

Eindrückung. Dies führt zu einer Erhöhung des e�ektiven Reibungskoef-

�zienten und zu einer nicht-trivialen Korrelation zwischen der Ritzhärte und

der Normalhärte von Gläsern.

Es wurde festgestellt, dass seitliche Eindrücke mit moderaten Belas-

tungskräften (bis zu 30 mN) auf der Ober�äche von Quarzglas ein gewisses

Ri�elphänomen bilden. Dies wurde bei verschiedenen Kratzgeschwindigkeiten

(bis zu 50 µm/s) weiter untersucht und wurde mit einer periodischen Stick-

Slip-Bewegung der Spitze in Verbindung gebracht. Die Periode der Rif-

felung wurde abgeschätzt und zeigte eine ziemlich lineare Abhängigkeit von

der Kratzgeschwindigkeit. Da die meisten technisch interessanten Gleit-

�ächen letztlich aus eine Vielzahl von Kontaktpunkten gebildet werden, die

abrupt gebildet und gebrochen werden, können die so gewonnenen Informa-

tionen von zentralem Interesse für das Verständnis viel komplexere Ver-

schleiÿprozesse auf der Makroskala. Konkret könnte man erwarten, dass

nicht nur der unregelmäÿige zeitliche Verlauf der Reibungskraft in einem

Gleitkontakt aus einem Ensemble von regelmäÿigen Stick-Slip-Ereignissen re-

sultiert, sondern die begleitenden abrasiven Verschleiÿmuster letztlich durch

Überlagerung (und mögliche Interferenz) von �Rripple-like �-Merkmalen verur-

sacht werden, ähnlich wie die hier für den eher idealen Fall einer umgekehrten

Pyramide vs bei Kontakt. Die Fourier-Analyse der während des Scannens

erfassten Querkraft kann auch helfen, die zeitliche Entwicklung des Stick-

Slip-Mechanismus mit den topographischen Eigenschaften der resultierenden

Ober�ächenmuster zu korrelieren.

Das Rampenbelastungskratzen wurde als geeignete Technik erkannt, um

die tribologischen Eigenschaften von Gläsern zu untersuchen. Da sich die

meisten Studien bisher darauf konzentrierten qualitative Informationen über

die Ober�äche von Gläsern zu erhalten, wurde die instrumentierte Nanoin-

dentation mit Rampenbelastung eingesetzt, um quantitative Informationen

über den Beginn des kratzinduzierten Mikroabriebs an Quarzglas zu erhal-

ten. Dazu wurde die in-situ-Auswertung der Querkraft und des Reibungsko-

e�zienten mit der mit der postmortalen optischen Inspektion nach dem Vor-

wärtskratzen mit einem Berkovich-Eindringkörper verglichen. Die statistis-

che Analyse ergab zwei zugrunde liegende Wahrscheinlichkeitsfunktionen für

Page 99: Tribological studies on the surface of glasses by lateral

das Auftreten von Mikroabrieb, d.h. die Wahrscheinlichkeit dass der sich

ausbreitende Kratzer eine Ober�ächenfehlstelle tri�t und die Wahrschein-

lichkeit, dass ein solches Ereignis einen beobachtbaren Mikroriss verursacht.

Die Dominanz des Erstere folgt einer Exponentialfunktion, die eine reine Zu-

fallsverteilung mit lastunabhängiger Ausfallwahrscheinlichkeit widerspiegelt.

Sie wurde nur bei hohen Kratzgeschwindigkeiten nach Überschreiten einer

bestimmten Normallast beobachtet. Für letztere wurde festgestellt, dass

der Weibull-Modul mit zunehmender Kratzgeschwindigkeit Geschwindigkeit,

d.h. von 1,6 auf 4,4. Hier wurde der niedrige Weibull-Modul bei niedriger

Last auf die zunehmende Zeit der lokalen Dehnung zurückgeführt, was zu

einer Verringerung der der Lastabhängigkeit der Mikrorissbildung führt. Für

Siliziumdioxid und unter Verwendung dem vorliegenden experimentellen Ansatz

liegt die kritische laterale Last für Mikroabrasion (50. Perzentil) bei etwa

30-40 mN. Dieser Wert ist sehr wahrscheinlich abhängig von extrinsischen

Parametern wie der Umgebungsfeuchte abhängig.

Des Weiteren wurde die Rampenlast-Ritztechnik verwendet, um tribolo-

gische Informationen über die Ober�äche des metallischen Glases (Zr55Cu30Al10Ni5)

zu erhalten, die ein groÿes Potential für industrielle Anwendungen hat. Auch

hier wurde bestätigt, dass das Kratzen ein stärkerer Verformungsmodus im

Vergleich zur normaler Eindrückung ist. Dies wurde durch die Untersuchung

der ersten pop-in auf P-h-Diagramme argumentiert. Der aus den Eindring-

daten geschätzte Reibungskoe�zient lag zwischen 0 und 0,15, also genau

zwischen den Reibungskoe�zienten von Gläsern und Metallen. Eine weit-

ere rasterelektronenmikroskopische Untersuchung der Kratzer auf der Ober-

�äche zeigte, dass es möglicherweise einen Zusammenhang zwischen dem Be-

ginn der Scherdeformation Verformung und den halbkreisförmigen Mustern,

die auf der Ober�äche beobachtet wurden.

Page 100: Tribological studies on the surface of glasses by lateral
Page 101: Tribological studies on the surface of glasses by lateral

Bibliography

[1] V. Le Houérou, J.-C. Sangleboeuf, S. Deriano, T. Rouxel, and

G. Duisit, �Surface damage of soda-lime-silica glasses: indentation

scratch behavior,� Journal of Non-Crystalline Solids, vol. 316, no. 1,

pp. PII S0022�3093(02)01937�3, 2003.

[2] F. Pariafsai, �A review of design considerations in glass buildings,�

Frontiers of Architectural Research, vol. 5, no. 2, pp. 171�193, 2016.

[3] S. Barbi, P. Miselli, and C. Siligardi, �Failure analysis of glazed

las glass-ceramic containing cerium oxide,� Ceramics International,

vol. 43, no. 1, Part B, pp. 1472�1478, 2017.

[4] C. Gao and M. Liu, �E�ects of normal load on the coe�cient of friction

by microscratch test of copper with a spherical indenter,� Tribology

letters, vol. 67, no. 1, pp. 1�12, 2018.

[5] P. Bandyopadhyay, A. Dey, and A. K. Mukhopadhyay, �Novel com-

bined scratch and nanoindentation experiments on soda-lime-silica

glass,� International Journal of Applied Glass Science, vol. 3, no. 2,

pp. 163�179, 2012.

[6] H. R. Chamani and M. R. Ayatollahi, �Prediction of friction coe�cients

in nanoscratch testing of metals based on material �ow lines,� Theo-

retical and Applied Fracture Mechanics, vol. 94, pp. 186�196, 2018.

[7] V. Le Houérou, J.-C. Sangleboeuf, and T. Rouxel, �Scratchability of

soda-lime silica (SLS) glasses: Dynamic fracture analysis,� Fractogra-

phy of Advanced Ceramics Ii, vol. 290, pp. 31�38, 2005.

[8] P. Bandyopadhyay, A. Dey, S. Roy, N. Dey, and A. K. Mukhopad-

hyay, �Nanomechanical properties inside the scratch grooves of soda-

87

Page 102: Tribological studies on the surface of glasses by lateral

88 Bibliography

lime-silica glass,� Applied Physics A: Materials Science & Processing,

vol. 107, pp. 943�948, June 2012.

[9] J. Schneider, S. Schula, and W. Weinhold, �Characterisation of the

scratch resistance of annealed and tempered architectural glass,� Thin

Solid Films, vol. 520, no. 12, pp. 4190 � 4198, 2012. 8th International

Conference on Coatings on Glass and Plastics ICCG8.

[10] K. H. Nielsen, S. Karlsson, R. Limbach, and L. Wondraczek, �Quantita-

tive image analysis for evaluating the abrasion resistance of nanoporous

silica �lms on glass,� Scienti�c Reports, vol. 5, 2015.

[11] C. M. Hermansen, J. Matsuoka, S. Yoshida, H. Yamazaki, Y. Kato, and

Y. Yue, �Densi�cation and plastic deformation under microindentation

in silicate glasses and the relation to hardness and crack resistance,�

Journal of Non-crystalline Solids, vol. 364, pp. 40�43, 2013.

[12] J.-C. Sangleb÷uf and T. Rouxel, �Indentation and scratching of glass:

Load, composition and temperature e�ects,� in Fracture Mechanics of

Ceramics (R. C. Bradt, D. Munz, M. Sakai, and K. W. White, eds.),

(Boston, MA), pp. 121�133, Springer US, 2005.

[13] K. Gahr, Microstructure and Wear of Materials. Tribology Series, El-

sevier, 1987.

[14] M. Wong, G. Lim, A. Moyse, J. Reddy, and H.-J. Sue, �A new test

methodology for evaluating scratch resistance of polymers,� Wear,

vol. 256, no. 11, pp. 1214 � 1227, 2004.

[15] J. Musgraves, J. Hu, and L. Calvez, Springer Handbook of Glass.

Springer Handbooks, Springer International Publishing, 2019.

[16] P. Burnett and D. Rickerby, �The relationship between hardness and

scratch adhession,� Thin Solid Films, vol. 154, no. 1, pp. 403�416,

1987.

[17] H. Ollendorf and D. Schneider, �A comparative study of adhesion test

methods for hard coatings,� Surface and Coatings Technology, vol. 113,

no. 1, pp. 86�102, 1999.

Page 103: Tribological studies on the surface of glasses by lateral

Bibliography 89

[18] D. Beegan, S. Chowdhury, and M. Laugier, �Comparison between

nanoindentation and scratch test hardness (scratch hardness) values

of copper thin �lms on oxidised silicon substrates,� Surface and Coat-

ings Technology, vol. 201, no. 12, pp. 5804�5808, 2007.

[19] F. Wredenberg and P.-L. Larsson, �Scratch testing of metals and poly-

mers: Experiments and numerics,� Wear, vol. 266, no. 1, pp. 76�83,

2009.

[20] A.-T. Akono, N. X. Randall, and F.-J. Ulm, �Experimental determi-

nation of the fracture toughness via microscratch tests: Application

to polymers, ceramics, and metals,� Journal of Materials Research,

vol. 27, no. 2, p. 485?493, 2012.

[21] R. Bard and F.-J. Ulm, �Scratch hardness-strength solutions for

cohesive-frictional materials,� International Journal for Numerical and

Analytical Methods in Geomechanics, vol. 36, no. 3, pp. 307�326, 2012.

[22] S. Miyake and S. Yamazaki, �Nanoscratch properties of extremely thin

diamond-like carbon �lms,� Wear, vol. 305, no. 1, pp. 69�77, 2013.

[23] Z. Wang, Q. Zeng, and J. Zheng, �Adsorption and lubricating behavior

of salivary pellicle on dental ceramic,� Lubrication Engineering, 2017.

[24] J. Fu, H. He, W. Yuan, Y. Zhang, and J. Yu, �Towards a deeper

understanding of nanoscratch-induced deformation in an optical glass,�

Applied Physics Letters, vol. 113, no. 3, 2018. cited By 8.

[25] R. F. Cook and G. M. Pharr, �Direct observation and analysis of in-

dentation cracking on glasses and ceramics,� American Ceramic Society

Bulletin; (USA), vol. 73:4, 4 1990.

[26] V. Howes and A. Szameitat, �Morphology of sub-surface cracks below

the scratch produced on soda lime glass by a disc cutter,� Journal of

Materials Science Letters, vol. 3, no. 10, pp. 872�874, 1984. cited By

5.

[27] P. Bandyopadhyay and A. K. Mukhopadhyay, �Role of shear stress

in scratch deformation of soda-lime-silica glass,� Journal of Non-

Crystalline Solids, vol. 362, pp. 101�113, 2013.

Page 104: Tribological studies on the surface of glasses by lateral

90 Bibliography

[28] B. Taljat and G. Pharr, �Development of pile-up during spherical in-

dentation of elastic?plastic solids,� International Journal of Solids and

Structures, vol. 41, no. 14, pp. 3891�3904, 2004.

[29] E. Gnecco, P. Pedraz, P. Nita, F. Dinelli, S. Napolitano, and P. Pingue,

�Surface rippling induced by periodic instabilities on a polymer sur-

face,� New Journal of Physics, vol. 17, p. 032001, mar 2015.

[30] O. Benzine, S. Bruns, Z. Pan, K. Durst, and L. Wondraczek, �Local

deformation of glasses is mediated by rigidity �uctuation on nanometer

scale,� Advanced Science, vol. 5, no. 10, p. 1800916, 2018.

[31] M. M. Trexler and N. N. Thadhani, �Mechanical properties of bulk

metallic glasses,� Progress in Materials Science, vol. 55, no. 8, pp. 759�

839, 2010.

[32] I.-C. Choi, Y. Zhao, B.-G. Yoo, Y.-J. Kim, J.-Y. Suh, U. Ramamurty,

and J. il Jang, �Estimation of the shear transformation zone size in

a bulk metallic glass through statistical analysis of the �rst pop-in

stresses during spherical nanoindentation,� Scripta Materialia, vol. 66,

no. 11, pp. 923 � 926, 2012. Viewpoint set no. 49: Strengthening e�ect

of nano-scale twins.

[33] D. Lahiri, J. Karp, A. K. Keshri, C. Zhang, G. S. Dulikravich, L. J.

Kecskes, and A. Agarwal, �Scratch induced deformation behavior of

hafnium based bulk metallic glass at multiple load scales,� Journal of

Non-Crystalline Solids, vol. 410, pp. 118 � 126, 2015.

[34] R. Limbach, K. Kosiba, S. Pauly, U. Kühn, and L. Wondraczek, �Ser-

rated �ow of cuzr-based bulk metallic glasses probed by nanoindenta-

tion: Role of the activation barrier, size and distribution of shear trans-

formation zones,� Journal of Non-Crystalline Solids, vol. 459, pp. 130�

141, 2017.

[35] D. Louzguine-Luzgin, S. Ketov, A. Trifonov, and A. Churymov, �Sur-

face structure and properties of metallic glasses,� Journal of Alloys and

Compounds, vol. 742, pp. 512�517, 2018.

[36] Y. Zhao, Y. Ye, C. Liu, R. Feng, K. Yao, and T. Nieh, �Tribolog-

ical behavior of an amorphous Zr20Ti20Cu20Ni20Be20 high-entropy

Page 105: Tribological studies on the surface of glasses by lateral

Bibliography 91

alloy studied using a nanoscratch technique,� Intermetallics, vol. 113,

p. 106561, 2019.

[37] P. Gong, F. Li, L. Deng, X. Wang, and J. Jin, �Research on nano-

scratching behavior of TiZrHfBeCu(Ni) high entropy bulk metallic

glasses,� Journal of Alloys and Compounds, vol. 817, p. 153240, 2020.

[38] S. Freiman, �Chapter 2 - fracture mechanics of glass,� in Elasticity and

Strength in Glasses (D. Uhlmann and N. Kreidl, eds.), vol. 5 of Glass

Science and Technology, pp. 21�78, Elsevier, 1980.

[39] G. I. Finch, A. G. Quarrell, J. S. Roebuck, and W. A. Bone, �The

beilby layer,� Proceedings of the Royal Society of London. Series A,

Containing Papers of a Mathematical and Physical Character, vol. 145,

no. 855, pp. 676�681, 1934.

[40] G. T. Beilby, �Researches on polished surfaces,� Transactions of the

Optical Society, vol. 9, pp. 22�71, jan 1907.

[41] E. W. Taylor, �A hardness table for some well-known types of optical

glass,� Journal of Scienti�c Instruments, vol. 26, pp. 314�316, sep 1949.

[42] P. W. Bridgman and I. Simon, �E�ects of very high pressures on glass,�

Journal of Applied Physics, vol. 24, no. 4, pp. 405�413, 1953.

[43] F. M. Ernsberger, �Role of densi�cation in deformation of glasses un-

der point loading,� Journal of the American Ceramic Society, vol. 51,

no. 10, pp. 545�547, 1968.

[44] B. Bhushan, Principles and Applications of Tribology. Tribology in

Practice Series, Wiley, 2013.

[45] A. Varshneya, Fundamentals of Inorganic Glasses. Elsevier Science,

1994.

[46] H. Hertz, �On the contact of elastic solids,� Z. Reine Angew. Mathe-

matik, vol. 92, pp. 156�171, 1881.

[47] B. Lawn and M. Swain, �Microfracture beneath point indenters in brit-

tle solids,� Journal of Materials Science, vol. 10, pp. 113�122, 07 1975.

Page 106: Tribological studies on the surface of glasses by lateral

92 Bibliography

[48] K. Nlihara, R. Morena, and D. P. H. Hasselman, �Further reply to

comment on elastic/plastic indentation damage in ceramics: The me-

dian/radial crack system,� Journal of the American Ceramic Society,

vol. 65, no. 7, pp. c116�c116, 1982.

[49] Y. Ahn, T. Farris, and S. Chandrasekar, �Sliding microindentation

fracture of brittle materials: Role of elastic stress �elds,� Mechanics of

Materials, vol. 29, no. 3, pp. 143�152, 1998.

[50] P. Bandyopadhyay, A. Dey, A. Mandal, N. Dey, S. Roy, and

A. Mukhopadhyay, �E�ect of scratching speed on deformation of soda-

lime-silica glass,� Applied Physics A, vol. 107, 06 2012.

[51] X. Jing, S. Maiti, and G. Subhash, �A new analytical model for es-

timation of scratch-induced damage in brittle solids,� Journal of the

American Ceramic Society, vol. 90, no. 3, pp. 885�892, 2007.

[52] J. D. B. Veldkamp, N. Hattu, and V. A. C. Snijders, Crack Forma-

tion during Scratching of Brittle Materials, pp. 273�301. Boston, MA:

Springer US, 1978.

[53] H. Ben Abdelounis, K. Elleuch, R. Vargiolu, H. Zahouani, and A. Le

Bot, �On the behaviour of obsidian under scratch test,� Wear, vol. 266,

no. 7, pp. 621�626, 2009.

[54] T. Gross, �Deformation and cracking behavior of glasses indented with

diamond tips of various sharpness,� Journal of Non-Crystalline Solids,

vol. 358, no. 24, pp. 3445�3452, 2012. Glasses for Energy.

[55] D. Marshall and B. Lawn, �Residual stress e�ects in sharp contact

cracking,� Journal of Materials Science, vol. 14, pp. 2001�2012, 1979.

[56] B. Lawn, Fracture of Brittle Solids. Cambridge Solid State Science

Series, Cambridge University Press, 2 ed., 1993.

[57] S. E. Flores, M. G. Pontin, and F. W. Zok, �Scratching of Elas-

tic/Plastic Materials With Hard Spherical Indenters,� Journal of Ap-

plied Mechanics, vol. 75, 08 2008. 061021.

[58] B. R. Lawn and A. G. Evans, �A model for crack initiation in elas-

tic/plastic indentation �elds,� Journal of Materials Science, vol. 12,

pp. 2195�2199, Nov. 1977.

Page 107: Tribological studies on the surface of glasses by lateral

Bibliography 93

[59] K. Peter, �Glastechn. bet. 37, 333. peter, k. & dick,� E.(1967).

Glastechn. Ber, vol. 40, no. 470, p. 28, 1964.

[60] D. Marshall, B. Lawn, and A. Evans, �Elastic/plastic indentation dam-

age in ceramics: The lateral crack system,� Journal of the American

Ceramic Society, vol. 65, no. 11, pp. 561�566, 1982.

[61] J.-P. Guin, T. Rouxel, J.-C. Sangleboeuf, I. Melscoët, and J. Lu-

cas, �Hardness, toughness, and scratchability of germanium-selenium

chalcogenide glasses,� Journal of the American Ceramic Society,

vol. 85, no. 6, pp. 1545�1552, 2002.

[62] F. Petit, C. Ott, and F. Cambier, �Multiple scratch tests and surface-

related fatigue properties of monolithic ceramics and soda lime glass,�

Journal of the European Ceramic Society, vol. 29, no. 8, pp. 1299�1307,

2009.

[63] W. Gu and Z. Yao, �Evaluation of surface cracking in micron and sub-

micron scale scratch tests for optical glass bk7,� Journal of Mechanical

Science and Technology, vol. 25, pp. 1167�1174, 05 2011.

[64] O. Borrero-López, M. Ho�man, A. Bendavid, and P. J. Martin, �Me-

chanical properties and scratch resistance of �ltered-arc-deposited ti-

tanium oxide thin �lms on glass,� Thin Solid Films, vol. 519, no. 22,

pp. 7925�7931, 2011.

[65] R. Crombez, J. McMinis, V. Veerasamy, and W. Shen, �Experimen-

tal study of mechanical properties and scratch resistance of ultra-thin

diamond-like-carbon (dlc) coatings deposited on glass,� Tribology In-

ternational, vol. 44, no. 1, pp. 55�62, 2011.

[66] F. Auerbach, �Über die härte-und elastizitätsverhaltnisse des glases,�

Wiedem. Annal, vol. 53, p. 1000, 1894.

[67] E. Brown Jr, �The friction and wear of glass,� ASLE Transactions,

vol. 12, no. 4, pp. 227�233, 1969.

[68] N. Bensaid, S. Benbahouche, F. Roumili, J.-C. Sangleboeuf, J.-B. L.

Cam, and T. Rouxel, �In�uence of the normal load of scratching on

cracking and mechanical strength of soda-lime-silica glass,� Journal of

Non-Crystalline Solids, vol. 483, pp. 65�69, 2018.

Page 108: Tribological studies on the surface of glasses by lateral

94 Bibliography

[69] B. R. Lawn and F. C. Frank, �Partial cone crack formation in a brit-

tle material loaded with a sliding spherical indenter,� Proceedings of

the Royal Society of London. Series A. Mathematical and Physical Sci-

ences, vol. 299, no. 1458, pp. 307�316, 1967.

[70] F. C. Frank and B. R. Lawn, �On the theory of hertzian fracture,�

Proceedings of the Royal Society of London. Series A. Mathematical

and Physical Sciences, vol. 299, no. 1458, pp. 291�306, 1967.

[71] Y. Wang and S. M. Hsu, �Wear and wear transition modeling of ce-

ramics,� Wear, vol. 195, no. 1, pp. 35�46, 1996.

[72] S. Yoshida, S. Iwata, T. Sugawara, Y. Miura, J. Matsuoka, A. Erra-

part, and C. R. Kurkjian, �Elastic and residual stresses around ball

indentations on glasses using a micro-photoelastic technique,� Jour-

nal of Non-Crystalline Solids, vol. 358, no. 24, pp. 3465�3472, 2012.

Glasses for Energy.

[73] K. Li, Y. Shapiro, and J. Li, �Scratch test of soda-lime glass,� Acta

Materialia, vol. 46, no. 15, pp. 5569 � 5578, 1998.

[74] M. Klecka and G. Subhash, �Grain size dependence of scratch-induced

damage in alumina ceramics,� Wear, vol. 265, no. 5, pp. 612�619, 2008.

[75] H. Kasem, S. Bonnamy, Y. Berthier, and P. Jacquemard, �Characteri-

zation of surface grooves and scratches induced by friction of c/c com-

posites at low and high temperatures,� Tribology International, vol. 43,

no. 11, pp. 1951�1959, 2010.

[76] P. Bandyopadhyay, A. Dey, A. K. Mandal, N. Dey, and A. K.

Mukhopadhyay, �New observations on scratch deformations of soda

lime silica glass,� Journal of Non-Crystalline Solids, vol. 358, no. 16,

pp. 1897�1907, 2012.

[77] N. J. Kreidl, �Glass surfaces then and now,� Journal of Non-Crystalline

Solids, vol. 49, no. 1, pp. 331�338, 1982. Proceedings of the Sixth

University Conference on Glass Science.

[78] S. M. Wiederhorn, J. M. López-Cepero, J. Wallace, J.-P. Guin, and

T. Fett, �Roughness of glass surfaces formed by sub-critical crack

Page 109: Tribological studies on the surface of glasses by lateral

Bibliography 95

growth,� Journal of Non-Crystalline Solids, vol. 353, no. 16, pp. 1582�

1591, 2007.

[79] A. Varshneya and J. Mauro, Fundamentals of Inorganic Glasses. El-

sevier Science, 2019.

[80] D. Uhlmann and N. Kreidl, Elasticity and Strength in Glasses. Glass�

science and Technology, Academic Press, 1980.

[81] P. Bandyopadhyay, A. Dey, S. Roy, and A. K. Mukhopadhyay, �E�ect

of load in scratch experiments on soda lime silica glass,� Journal of

Non-Crystalline Solids, vol. 358, no. 8, pp. 1091�1103, 2012.

[82] S. Yoshida, T. Hayashi, T. Fukuhara, K. Soeda, J. Matsuoka, and

N. Soga, Scratch Test for Evaluation of Surface Damage in Glass,

pp. 101�111. 01 2005.

[83] S. Wiederhorn, S. Freimann, J. Fuller, and C. Simmons, �E�ect of

water and other dielectrics on crack growth,� Journal of Materials Sci-

ence, vol. 17, p. 3460?3478, 1982.

[84] S. N. Crichton, M. Tomozawa, J. S. Hayden, T. I. Suratwala, and

J. H. Campbell, �Subcritical crack growth in a phosphate laser glass,�

Journal of the American Ceramic Society, vol. 82, no. 11, pp. 3097�

3104, 1999.

[85] C. L. Rountree, �Recent progress to understand stress corrosion crack-

ing in sodium borosilicate glasses: linking the chemical composition

to structural, physical and fracture properties,� Journal of Physics D:

Applied Physics, vol. 50, p. 343002, aug 2017.

[86] S. Yoshida, J. Matsuoka, and N. Soga, �Sub-critical crack growth in

sodium germanate glasses,� Journal of Non-Crystalline Solids, vol. 316,

no. 1, pp. 28�34, 2003.

[87] T. A. Michalske and B. C. Bunker, �Slow fracture model based on

strained silicate structures,� Journal of Applied Physics, vol. 56, no. 10,

pp. 2686�2693, 1984.

[88] E. Silva, J. Li, D. Liao, S. Subramanian, T. Zhu, and S. Yip, �Atomic

scale chemo-mechanics of silica: nano-rod deformation and water re-

Page 110: Tribological studies on the surface of glasses by lateral

96 Bibliography

action,� Journal of computer-aided materials design, vol. 13, no. 1-3,

pp. 135�159, 2006.

[89] T. Zhu, J. Li, X. Lin, and S. Yip, �Stress-dependent molecular path-

ways of silica-water reaction,� Journal of the Mechanics and Physics

of Solids, vol. 53, no. 7, pp. 1597�1623, 2005.

[90] Y.-F. Niu, K. Han, and J.-P. Guin, �Locally enhanced dissolution rate

as a probe for nanocontact-induced densi�cation in oxide glasses,�

Langmuir, vol. 28, no. 29, pp. 10733�10740, 2012.

[91] B. A. Proctor, I. Whitney, J. W. Johnson, and A. H. Cottrell, �The

strength of fused silica,� Proceedings of the Royal Society of London.

Series A. Mathematical and Physical Sciences, vol. 297, no. 1451,

pp. 534�557, 1967.

[92] V. Houérou, Rayabilité des verres Silico-Sodo-Calciques. PhD thesis,

Université de Rennes 1, 07 2005.

[93] S. M. Wiederhorn, �In�uence of water vapor on crack propagation in

soda-lime glass,� Journal of the American Ceramic Society, vol. 50,

no. 8, pp. 407�414, 1967.

[94] R. E. Mould, �Strength and static fatigue of abraded glass under con-

trolled ambient conditions: Iv, e�ect of surrounding medium,� Journal

of the American Ceramic Society, vol. 44, no. 10, pp. 481�491, 1961.

[95] M. Muraoka and H. Abé, �Subcritical crack growth in silica optical

�bers in wide range of crack velocities,� Journal of the American Ce-

ramic Society, vol. 79, no. 1, pp. 51�57, 1996.

[96] M. Chaudhri and C. Kurkjian, �Impact of small steel spheres on the

surfaces of normal and anomalous glasses,� Journal of the American

Ceramic Society, vol. 69, no. 5, pp. 404�410, 1986.

[97] G. Greaves, �EXAFS and the structure of glass,� Journal of Non-

Crystalline Solids, vol. 71, no. 1, pp. 203�217, 1985. E�ects of Modes

of Formation on the Structure of Glass.

[98] G. N. de Macedo, S. Sawamura, and L. Wondraczek, �Lateral hardness

and the scratch resistance of glasses in the Na2O-CaO-SiO2 system,�

Journal of Non-Crystalline Solids, vol. 492, pp. 94�101, 2018.

Page 111: Tribological studies on the surface of glasses by lateral

Bibliography 97

[99] C. Meade, R. J. Hemley, and H. K. Mao, �High-pressure x-ray di�rac-

tion of SiO2 glass,� Phys. Rev. Lett., vol. 69, pp. 1387�1390, Aug 1992.

[100] A. Arora, D. Marshall, B. Lawn, and M. Swain, �Indentation defor-

mation/fracture of normal and anomalous glasses,� Journal of Non-

crystalline Solids, vol. 31, pp. 415�428, 1979.

[101] T. Rouxel, H. Ji, J. P. Guin, F. Augereau, and B. Ru�é, �Indenta-

tion deformation mechanism in glass: Densi�cation versus shear �ow,�

Journal of Applied Physics, vol. 107, no. 9, p. 094903, 2010.

[102] K. Peter, �Densi�cation and �ow phenomena of glass in indentation

experiments,� Journal of Non Crystalline Solids, vol. 5, pp. 103�115,

Nov. 1970.

[103] T. Rouxel, H. Ji, T. Hammouda, and A. Moréac, �Poisson's ratio

and the densi�cation of glass under high pressure,� Phys. Rev. Lett.,

vol. 100, p. 225501, Jun 2008.

[104] J. Sehgal and S. Ito, �Brittleness of glass,� Journal of Non-Crystalline

Solids, vol. 253, no. 1, pp. 126 � 132, 1999.

[105] L. Wondraczek, J. C. Mauro, J. Eckert, U. Kühn, J. Horbach,

J. Deubener, and T. Rouxel, �Towards ultrastrong glasses,� Advanced

Materials, vol. 23, no. 39, pp. 4578�4586, 2011.

[106] B. R. Lawn and D. B. Marshall, �Hardness, toughness, and brittleness:

An indentation analysis,� Journal of the American Ceramic Society,

vol. 62, no. 7?8, pp. 347�350, 1979.

[107] J. D. Mackenzie, �High-pressure e�ects on oxide glasses: I, densi�ca-

tion in rigid state,� Journal of the American Ceramic Society, vol. 46,

no. 10, pp. 461�470, 1963.

[108] S. Yoshida, J.-C. Sangleboeuf, and T. Rouxel, �Quantitative evaluation

of indentation-induced densi�cation in glass,� Journal of Materials Re-

search, vol. 20, no. 12, p. 3404?3412, 2005.

[109] S. Yoshida, S. Isono, J. Matsuoka, and N. Soga, �Shrinkage behavior

of knoop indentations in silica and soda-lime-silica glasses,� Journal of

the American Ceramic Society, vol. 84, no. 9, pp. 2141�2143, 2001.

Page 112: Tribological studies on the surface of glasses by lateral

98 Bibliography

[110] J. Kjeldsen, M. M. Smedskjaer, J. C. Mauro, and Y. Yue, �Hardness

and incipient plasticity in silicate glasses: Origin of the mixed modi�er

e�ect,� Applied Physics Letters, vol. 104, no. 5, p. 051913, 2014.

[111] J. Kjeldsen, M. M. Smedskjaer, J. C. Mauro, and Y. Yue, �On the origin

of the mixed alkali e�ect on indentation in silicate glasses,� Journal of

Non-Crystalline Solids, vol. 406, pp. 22�26, 2014.

[112] R. Limbach, A. Winterstein-Beckmann, J. Dellith, D. Möncke, and

L. Wondraczek, �Plasticity, crack initiation and defect resistance in

alkali-borosilicate glasses: From normal to anomalous behavior,� Jour-

nal of Non-Crystalline Solids, vol. 417-418, pp. 15 � 27, 2015.

[113] T. Rouxel, �Elastic properties and short-to medium-range order in

glasses,� Journal of the American Ceramic Society, vol. 90, no. 10,

pp. 3019�3039, 2007.

[114] T. Deschamps, A. Kassir-Bodon, C. Sonneville, J. Margueritat,

C. Martinet, D. de Ligny, A. Mermet, and B. Champagnon, �Perma-

nent densi�cation of compressed silica glass: a raman-density calibra-

tion curve,� Journal of Physics: Condensed Matter, vol. 25, p. 025402,

nov 2012.

[115] G. N. Greaves, A. L. Greer, R. S. Lakes, and T. Rouxel, �Poisson's ratio

and modern materials,� Nature Materials, vol. 10, no. 11, pp. 823�837,

2011.

[116] A. T. AlMotasem, J. Bergström, A. Gåård, P. Krakhmalev, and

L. Holleboom, �Atomistic insights on the wear/friction behavior of

nanocrystalline ferrite during nanoscratching as revealed by molecu-

lar dynamics,� Tribology letters, vol. 65, no. 3, pp. 1�13, 2017.

[117] A. Diez-Ibarbia, A. Fernandez-Del-Rincon, P. Garcia, A. De-Juan,

M. Iglesias, and F. Viadero, �Assessment of load dependent friction

coe�cients and their in�uence on spur gears e�ciency,� Meccanica,

vol. 53, no. 1, pp. 425�445, 2018.

[118] O. Sterner, R. Aeschlimann, S. Zürcher, C. Scales, D. Riederer, N. D.

Spencer, and S. Tosatti, �Tribological classi�cation of contact lenses:

from coe�cient of friction to sliding work,� Tribology Letters, vol. 63,

no. 1, p. 9, 2016.

Page 113: Tribological studies on the surface of glasses by lateral

Bibliography 99

[119] S. Zhang, X. Zeng, A. Igartua, E. Rodriguez-Vidal, and E. Van der

Heide, �Texture design for reducing tactile friction independent of slid-

ing orientation on stainless steel sheet,� Tribology letters, vol. 65, no. 3,

p. 89, 2017.

[120] S. Li, Q. Li, R. Carpick, P. Gumbsch, X. Liu, X. Ding, L. Juan, and

J. Li, �The evolving quality of frictional contact with graphene,� Na-

ture, vol. 539, pp. 541�545, 11 2016.

[121] P. Saravanan, R. Selyanchyn, M. Watanabe, S. Fujikawa, H. Tanaka,

S. M. Lyth, and J. Sugimura, �Ultra-low friction of polyethylenimine

/ molybdenum disul�de (PEI/MoS2)15 thin �lms in dry nitrogen at-

mosphere and the e�ect of heat treatment,� Tribology International,

vol. 127, pp. 255�263, 2018.

[122] P. Gabriel, A. Thomas, and J. Bus�eld, �In�uence of interface geometry

on rubber friction,� Wear, vol. 268, no. 5, pp. 747�750, 2010.

[123] S. Maegawa, F. Itoigawa, and T. Nakamura, �E�ect of normal load

on friction coe�cient for sliding contact between rough rubber surface

and rigid smooth plane,� Tribology International, vol. 92, pp. 335�343,

2015.

[124] T. Yamaguchi, T. Sugawara, M. Takahashi, K. Shibata, K. Moriyasu,

T. Nishiwaki, and K. Hokkirigawa, �E�ect of porosity and normal

load on dry sliding friction of polymer foam blocks,� Tribology Let-

ters, vol. 66, no. 1, pp. 1�9, 2018.

[125] V. Westlund, J. Heinrichs, and S. Jacobson, �On the role of material

transfer in friction between metals: initial phenomena and e�ects of

roughness and boundary lubrication in sliding between aluminium and

tool steels,� Tribology Letters, vol. 66, no. 3, pp. 1�15, 2018.

[126] P. Lu, R. J. Wood, M. G. Gee, L. Wang, and W. P�eging, �A novel

surface texture shape for directional friction control,� Tribology Letters,

vol. 66, no. 1, pp. 1�13, 2018.

[127] I. Szlufarska, M. Chandross, and R. W. Carpick, �Recent advances in

single-asperity nanotribology,� Journal of Physics D: Applied Physics,

vol. 41, p. 123001, may 2008.

Page 114: Tribological studies on the surface of glasses by lateral

100 Bibliography

[128] P. Udaykant Jadav, R. Amali, and O. B. Adetoro, �Analytical fric-

tion model for sliding bodies with coupled longitudinal and transverse

vibration,� Tribology International, vol. 126, pp. 240�248, 2018.

[129] W. Hardy, �III. the spreading of �uids on glass,� The London, Ed-

inburgh, and Dublin Philosophical Magazine and Journal of Science,

vol. 38, no. 223, pp. 49�55, 1919.

[130] W. Hardy, �Some problems of lubrication 1,� Nature, vol. 106, pp. 569�

572, 1920.

[131] V. Frechette, I. Ceramics, W. LaCourse, and V. Burdick, Surfaces and

Interfaces of Glass and Ceramics. Materials Science Research, Springer

US, 1974.

[132] F. Bowden, F. Bowden, and D. Tabor, The Friction and Lubrication

of Solids. No. v. 1 in International series of monographs on physics,

Clarendon Press, 2001.

[133] M. Liu, Q. Zheng, and C. Gao, �Sliding of a diamond sphere on fused

silica under ramping load,� Materials Today Communications, vol. 25,

p. 101684, 2020.

[134] Q. Wei, W. Gao, Q. Yang, and X. Li, �Material removal and tribological

behaviors of fused silica scratched by rockwell indenters with di�erent

tip radii,� Journal of Non-Crystalline Solids, vol. 514, pp. 90 � 97,

2019.

[135] F. P. Bowden and L. Leben, �The nature of sliding and the analy-

sis of friction,� Proceedings of the Royal Society of London. Series A.

Mathematical and Physical Sciences, vol. 169, no. 938, pp. 371�391,

1939.

[136] C. Wang, S. Z. Bonyadi, F. Grün, G. Pinter, A. Hausberger, and A. C.

Dunn, �Precise correlation of contact area and forces in the unsta-

ble friction between a rough �uoroelastomer surface and borosilicate

glass,� Materials, vol. 13, no. 20, 2020.

[137] Z. Elkaakour, J. P. Aimé, T. Bouhacina, C. Odin, and T. Masuda,

�Bundle formation of polymers with an atomic force microscope in

Page 115: Tribological studies on the surface of glasses by lateral

Bibliography 101

contract mode: A friction versus peeling process,� Phys. Rev. Lett.,

vol. 73, pp. 3231�3234, Dec 1994.

[138] R. H. Schmidt, G. Haugstad, and W. L. Gladfelter, �Correlation of

nanowear patterns to viscoelastic response in a thin polystyrene melt,�

Langmuir, vol. 15, no. 2, pp. 317�321, 1999.

[139] O. M. Leung and M. C. Goh, �Orientational ordering of polymers

by atomic force microscope tip-surface interaction,� Science, vol. 255,

no. 5040, pp. 64�66, 1992.

[140] B. Bhushan and A. V. Kulkarni, �E�ect of normal load on microscale

friction measurements,� Thin Solid Films, vol. 278, no. 1, pp. 49�56,

1996.

[141] C. Li, F. Zhang, Y. Ding, and L. Liu, �Surface deformation and friction

characteristic of nano scratch at ductile-removal regime for optical glass

bk7,� Appl. Opt., vol. 55, pp. 6547�6553, Aug 2016.

[142] H. He, S. H. Kim, and L. Qian, �E�ects of contact pressure, counter-

surface and humidity on wear of soda-lime-silica glass at nanoscale,�

Tribology International, vol. 94, pp. 675�681, 2016.

[143] C. M. Opri³an, B. Chiriac, V. Cârlescu, and D. N. Olaru, �In�u-

ence of the sti�ness and the speed on the stick-slip process,� vol. 997,

p. 012016, dec 2020.

[144] A. Fischer-Cripps, Nanoindentation. Mechanical Engineering Series,

Springer New York, 2011.

[145] M. Idriss, F. Célarié, Y. Yokoyama, F. Tessier, and T. Rouxel, �Evo-

lution of the elastic modulus of Zr-Cu-Al BMGs during annealing

treatment and crystallization: Role of Zr/Cu ratio,� Journal of Non-

Crystalline Solids, vol. 421, pp. 35 � 40, 2015.

[146] Y. Yokoyama, Y. Akeno, T. Yamasaki, P. K. Liaw, R. A.

Buchanan, and A. Inoue, �Evolution of mechanical properties of cast

Zr50Cu40Al10 glassy alloys by structural relaxation,� Materials Trans-

actions, vol. 46, no. 12, pp. 2755�2761, 2005.

Page 116: Tribological studies on the surface of glasses by lateral

102 Bibliography

[147] Y. Yokoyama, P. K. Liaw, M. Nishijima, K. Hiraga, R. A. Buchanan,

and A. Inoue, �Fatigue-strength enhancement of cast Zr50Cu40Al10

glassy alloys,� Materials Transactions, vol. 47, no. 5, pp. 1286�1293,

2006.

[148] Y. Yokoyama, T. Yamasaki, P. K. Liaw, and A. Inoue, �Study of the

structural relaxation-induced embrittlement of hypoeutectic zr-cu-al

ternary bulk glassy alloys,� Acta Materialia, vol. 56, no. 20, pp. 6097�

6108, 2008.

[149] I. Horcas, R. Fernández, J. M. Gómez-Rodríguez, J. Colchero,

J. Gómez-Herrero, and A. M. Baro, �Wsxm: A software for scanning

probe microscopy and a tool for nanotechnology,� Review of Scienti�c

Instruments, vol. 78, no. 1, p. 013705, 2007.

[150] S. Sawamura, R. Limbach, H. Behrens, and L. Wondraczek, �Lateral

deformation and defect resistance of compacted silica glass: Quanti�-

cation of the scratching hardness of brittle glasses,� Journal of Non-

Crystalline Solids, vol. 481, pp. 503 � 511, 2018.

[151] R. Limbach, B. P. Rodrigues, and L. Wondraczek, �Strain-rate sensi-

tivity of glasses,� Journal of Non-Crystalline Solids, vol. 404, pp. 124

� 134, 2014.

[152] A. Makishima and J. D. Mackenzie, �Calculation of bulk modulus,

shear modulus and poisson's ratio of glass,� Journal of Non-Crystalline

Solids, vol. 17, no. 2, pp. 147 � 157, 1975.

[153] S. Yoshida, Y. Hayashi, A. Konno, T. Sugawara, Y. Miura, and J. Mat-

suoka, �Indentation induced densi�cation of sodium borate glasses,�

Physics and Chemistry of Glasses - European Journal of Glass Science

and Technology Part B, vol. 50, pp. 63�70, 02 2009.

[154] M. Yamane and J. Mackenzie, �Vicker's hardness of glass,� Journal of

Non-Crystalline Solids, vol. 15, no. 2, pp. 153�164, 1974.

[155] J. Neely and J. Mackenzie, �Hardness and low-temperature deforma-

tion of silica glass,� Journal of Materials Science, vol. 3, no. 6, pp. 603�

609, 1968.

Page 117: Tribological studies on the surface of glasses by lateral

Bibliography 103

[156] H. Sawasato, S. Yoshida, T. Sugawara, Y. Miura, and J. Matsouka,

�Relaxation behaviors of vickers indentations in soda-lime glass,� Jour-

nal of the Ceramic Society of Japan, vol. 116, no. 1356, pp. 864�868,

2008.

[157] S. Yoshida, H. Sawasato, T. Sugawara, Y. Miura, and J. Matsuoka, �Ef-

fects of indenter geometry on indentation-induced densi�cation of soda-

lime glass,� Journal of Materials Research, vol. 25, no. 11, pp. 2203�

2211, 2010.

[158] E. Moayedi and L. Wondraczek, �Quantitative analysis of scratch-

induced microabrasion on silica glass,� Journal of Non-Crystalline

Solids, vol. 470, pp. 138 � 144, 2017.

[159] S. Yoshida, J.-C. Sangleboeuf, and T. Rouxel, �Indentation-induced

densi�cation of soda-lime silicate glass,� International Journal of Ma-

terials Research, vol. 98, no. 5, pp. 360�364, 2007.

[160] J.-F. Poggemann, G. Heide, and G. Frischat, �Direct view of the struc-

ture of di�erent glass fracture surfaces by atomic force microscopy,�

Journal of Non-Crystalline Solids, vol. 326-327, pp. 15�20, 2003. 13th

Int. Symp. on Non-Oxide Glasses and New Optical Materials.

[161] G. Frischat, J.-F. Poggemann, and G. Heide, �Nanostructure and

atomic structure of glass seen by atomic force microscopy,� Journal

of Non-Crystalline Solids, vol. 345-346, pp. 197�202, 2004. Physics of

Non-Crystalline Solids 10.

[162] D. A. Lind and S. P. Sanders, �Snow: The playing �eld,� in The Physics

of Skiing, pp. 11�43, Springer, 2004.

[163] J.-G. P. Tae-Young Kwon, Manivannan Ramachandran, �Scratch

formation and its mechanism in chemical mechanical planarization

(cmp),� Friction, vol. 1, no. 4, p. 279, 2013.

[164] H. Jiang, Q. Cheng, C. Jiang, J. Zhang, and L. Yonghua, �E�ect of

stick-slip on the scratch performance of polypropylene,� Tribology In-

ternational, vol. 91, pp. 1 � 5, 2015.

[165] A. Socoliuc, R. Bennewitz, E. Gnecco, and E. Meyer, �Transition from

stick-slip to continuous sliding in atomic friction: Entering a new

Page 118: Tribological studies on the surface of glasses by lateral

104 Bibliography

regime of ultralow friction,� Phys. Rev. Lett., vol. 92, p. 134301, Apr

2004.

[166] B. Such, F. Krok, and M. Szymonski, �AFM tip-induced ripple pattern

on AIII-BV semiconductor surfaces,� Applied Surface Science, vol. 254,

no. 17, pp. 5431�5434, 2008.

[167] B. Lawn, S. Wiederhorn, and D. Roberts, �E�ect of sliding friction

forces on the strength of brittle materials,� Journal of Materials Sci-

ence, vol. 19, pp. 2561�2569, 1984.

[168] S. Sawamura and L. Wondraczek, �Scratch hardness of glass,� Phys.

Rev. Materials, vol. 2, p. 092601, Sep 2018.

[169] T. Miura, Y. Benino, R. Sato, and T. Komatsu, �Universal hardness

and elastic recovery in vickers nanoindentation of copper phosphate

and silicate glasses,� Journal of the European Ceramic Society, vol. 23,

no. 3, pp. 409 � 416, 2003.

[170] P. Jutzi and U. Schubert, Silicon chemistry: from the atom to extended

systems. John Wiley & Sons, 2007.

[171] R. B. Jackman, J. S. Foord, A. E. Adams, and M. L. Lloyd, �Laser

chemical vapor deposition of patterned fe on silica glass: Observation

and origins of periodic ripple structures,� Journal of Applied Physics,

vol. 59, no. 6, pp. 2031�2034, 1986.

[172] C. Mate, Tribology on the Small Scale: A Bottom Up Approach to Fric-

tion, Lubrication, and Wear. Mesoscopic Physics and Nanotechnology,

OUP Oxford, 2008.

[173] R. Stapelberg, Handbook of Reliability, Availability, Maintainability

and Safety in Engineering Design. Springer-Verlag London, 2009.

[174] R. Abernethy, J. Breneman, C. Medlin, and G. Reinman, �Weibull

analysis handbook,� p. 243, 11 1983.

[175] R. H. Doremus, �Fracture statistics: A comparison of the normal,

weibull, and type I extreme value distributions,� Journal of Applied

Physics, vol. 54, no. 1, pp. 193�198, 1983.

Page 119: Tribological studies on the surface of glasses by lateral

Bibliography 105

[176] G. Quinn, �NIST recommended practice guide: Fractography of ce-

ramics and glasses,� 2006.

[177] J. Sehgal and S. Ito, �A new low-brittleness glass in the soda-lime-silica

glass family,� Journal of the American Ceramic Society, vol. 81, no. 9,

pp. 2485�2488, 1998.

[178] R. Mészáros, M. Wild, B. Merle, and L. Wondraczek, �Flexural

strength of PVD coated �oat glass for architectural applications,� 2011.

[179] F. Wredenberg and P.-L. Larsson, �On the numerics and correlation

of scratch testing,� Journal of Mechanics of Materials and Structures,

vol. 2, pp. 573�594, 03 2007.

[180] S. Bellemare, M. Dao, and S. Suresh, �The frictional sliding response

of elasto-plastic materials in contact with a conical indenter,� Interna-

tional Journal of Solids and Structures, vol. 44, no. 6, pp. 1970 � 1989,

2007. Physics and Mechanics of Advanced Materials.

[181] S. Bellemare, M. Dao, and S. Suresh, �E�ects of mechanical properties

and surface friction on elasto-plastic sliding contact,� Mechanics of

Materials, vol. 40, pp. 206�219, 04 2008.

[182] K. Johnson, Contact Mechanics. Cambridge University Press, 1985.

[183] L. Liu, L. Dai, Y. Bai, B. Wei, and J. Eckert, �Behavior of multiple

shear bands in Zr-based bulk metallic glass,� Materials Chemistry and

Physics, vol. 93, no. 1, pp. 174 � 177, 2005.

[184] W. Jiang, G. Fan, F. Liu, G. Wang, H. Choo, and P. Liaw, �Spatiotem-

porally inhomogeneous plastic �ow of a bulk-metallic glass,� Interna-

tional Journal of Plasticity, vol. 24, no. 1, pp. 1 � 16, 2008.

[185] S. Song and T. Nieh, �Flow serration and shear-band viscosity during

inhomogeneous deformation of a Zr-based bulk metallic glass,� Inter-

metallics, vol. 17, no. 9, pp. 762 � 767, 2009.

[186] B. A. Sun, H. B. Yu, W. Jiao, H. Y. Bai, D. Q. Zhao, and W. H. Wang,

�Plasticity of ductile metallic glasses: A self-organized critical state,�

Phys. Rev. Lett., vol. 105, p. 035501, Jul 2010.

Page 120: Tribological studies on the surface of glasses by lateral

106 Bibliography

[187] R. Maass, D. Klaumuenzer, and J. Loe�er, �Propagation dynamics of

individual shear bands during inhomogeneous �ow in a Zr-based bulk

metallic glass,� Acta Materialia, vol. 59, no. 8, pp. 3205�3213, 2011.

[188] C. A. Schuh and A. C. Lund, �Application of nucleation theory to

the rate dependence of incipient plasticity during nanoindentation,�

Journal of Materials Research, vol. 19, no. 7, p. 2152?2158, 2004.

[189] H. Bei, Z. P. Lu, and E. P. George, �Theoretical strength and the onset

of plasticity in bulk metallic glasses investigated by nanoindentation

with a spherical indenter,� Phys. Rev. Lett., vol. 93, p. 125504, Sep

2004.

[190] C. Schuh, J. Mason, and A. Lund, �Quantitative insight into disloca-

tion nucleation from high-temperature nanoindentation experiments,�

Nature materials, vol. 4, pp. 617�21, 09 2005.

[191] S. Shim, H. Bei, E. George, and G. Pharr, �A di�erent type of inden-

tation size e�ect,� Scripta Materialia, vol. 59, no. 10, pp. 1095 � 1098,

2008.

[192] C. Packard and C. Schuh, �Initiation of shear bands near a stress con-

centration in metallic glass,� Acta Materialia, vol. 55, no. 16, pp. 5348

� 5358, 2007.

[193] C. Pan, T. Wu, C. Liu, C. Su, W. Wang, and J. Huang, �Study of

scratching Mg-based BMG using nanoindenter with berkovich probe,�

Materials Science and Engineering: A, vol. 527, no. 9, pp. 2342 � 2349,

2010.

[194] B. R. Sun and Z. J. Zhan, �Indentation and friction of Cu44Zr43Al10Nb3bulk metallic glasses on nano-scale,� in Mechanical and Electronics

Engineering III, vol. 130 of Applied Mechanics and Materials, pp. 2750�

2753, Trans Tech Publications Ltd, 1 2012.

[195] S. Yoon, C. Lee, and H. Choi, �Evaluation of the e�ects of the crys-

tallinity of kinetically sprayed Ni-Ti-Zr-Si-Sn bulk metallic glass on the

scratch response,� Materials Science and Engineering: A, vol. 449-451,

pp. 285 � 289, 2007. Proceedings of the 12th International Conference

on Rapidly Quenched and Metastable Materials.

Page 121: Tribological studies on the surface of glasses by lateral

Bibliography 107

[196] A. Greer, Y. Cheng, and E. Ma, �Shear bands in metallic glasses,�

Materials Science and Engineering: R: Reports, vol. 74, no. 4, pp. 71

� 132, 2013.

[197] I.-C. Choi, Y. Zhao, Y.-J. Kim, B.-G. Yoo, J.-Y. Suh, U. Ramamurty,

and J. il Jang, �Indentation size e�ect and shear transformation zone

size in a bulk metallic glass in two di�erent structural states,� Acta

Materialia, vol. 60, no. 19, pp. 6862 � 6868, 2012.

[198] X. Zhao, Q. Cao, C. Wang, X. Wang, D. Zhang, S. Qu, and J. Jiang,

�Dependence of room-temperature nanoindentation creep behavior and

shear transformation zone on the glass transition temperature in

bulk metallic glasses,� Journal of Non-Crystalline Solids, vol. 445-446,

pp. 19 � 29, 2016.

[199] Y. Zhao, I.-C. Choi, M.-Y. Seok, M.-H. Kim, D.-H. Kim, U. Rama-

murty, J.-Y. Suh, and J. il Jang, �E�ect of hydrogen on the yielding

behavior and shear transformation zone volume in metallic glass rib-

bons,� Acta Materialia, vol. 78, pp. 213 � 221, 2014.

[200] J. H. Perepezko, S. D. Imho�, M.-W. Chen, J.-Q. Wang, and S. Gon-

zalez, �Nucleation of shear bands in amorphous alloys,� Proceedings

of the National Academy of Sciences, vol. 111, no. 11, pp. 3938�3942,

2014.

[201] W. Li, Y. Gao, and H. Bei, �On the correlation between micro-

scopic structural heterogeneity and embrittlement behavior in metallic

glasses,� Scienti�c Reports, vol. 5, p. 14786, 10 2015.

[202] C. Packard, O. Franke, E. Homer, and C. Schuh, �Nanoscale strength

distribution in amorphous versus crystalline metals,� Journal of Ma-

terials Research, vol. 25, no. 12, pp. 2251�2263, 2010.

[203] D. Tönnies, K. Samwer, P. M. Derlet, C. A. Volkert, and R. Maass,

�Rate-dependent shear-band initiation in a metallic glass,� Applied

Physics Letters, vol. 106, no. 17, p. 171907, 2015.

[204] W. L. Johnson and K. Samwer, �A universal criterion for plastic yield-

ing of metallic glasses with a (t/Tg)2/3 temperature dependence,� Phys.

Rev. Lett., vol. 95, p. 195501, Nov 2005.

Page 122: Tribological studies on the surface of glasses by lateral

108 Bibliography

[205] W. Wang, �Elastic moduli and behaviors of metallic glasses,� Journal

of Non-Crystalline Solids, vol. 351, no. 16, pp. 1481 � 1485, 2005.

[206] V. Keryvin, R. Crosnier, R. Laniel, V. H. Hoang, and J.-C. Sangleb÷uf,

�Indentation and scratching mechanisms of a ZrCuAlNi bulk metallic

glass,� Journal of Physics D: Applied Physics, vol. 41, p. 074029, mar

2008.

[207] M. Liu, J. Wu, and C. Gao, �Sliding of a diamond sphere on K9 glass

under progressive load,� Journal of Non-Crystalline Solids, vol. 526,

p. 119711, 2019.

[208] S. Leroch, M. Varga, S. Eder, A. Vernes, M. Rodriguez Ripoll, and

G. Ganzenmüller, �Smooth particle hydrodynamics simulation of dam-

age induced by a spherical indenter scratching a viscoplastic material,�

International Journal of Solids and Structures, vol. 81, pp. 188 � 202,

2016.

Page 123: Tribological studies on the surface of glasses by lateral

Appendix

Derivation of equation 4.2 in the main text In the most basic approx-

imation the nanoindenter tip can be modelled as a rigid object of mass m

which is pulled laterally (with constant velocity v) by a spring of sti�ness k.

If Fs is the static friction force, the tip will not move till the time

t1 =Fs

kv(E1)

At this point, the tip will respond as a harmonic oscillator subjected to

a negative (kinetic) friction force Fk (< Fs) slowing it down. The equation

of motion can be easily solved giving

x(t) = vt− Fk

k− a sin(ω0t+ ϕ) (E2)

for the tip position along the surface. In eq. E2 ω0 = (k/m)1/2 is the

resonance frequency of the spring+tip system. The oscillation amplitude a

and the phase shift ϕ are given by

a = (v2

ω2+

(Fs − Fk)2

k2)1/2 (E3)

109

Page 124: Tribological studies on the surface of glasses by lateral

110 Appendix

Figure F1: (a) Width and (b) depth of the wear grooves formed on a silica

glass surface scratched by a Berkovich diamond tip with a scan velocity of

10 µm/s and di�erent normal loads.

Figure F2: (a) AFM topography corresponding to Fig. 4.10(a); (b) horizon-

tal cross-section along the blue line in (a).

Page 125: Tribological studies on the surface of glasses by lateral

Appendix 111

Figure F3: AFM topography of the very end of the scratch. Frame size: 14.1

µm∗6.1 µm.

Figure F4: (a) AFM topography of one section of scratch performed at load

of 30 mN. (b) Pro�le along the scratch shown with a horizontal line in (a).

(c) FFT analysis of the same image. (d) Pro�le along the line in (c).

Page 126: Tribological studies on the surface of glasses by lateral

112

Page 127: Tribological studies on the surface of glasses by lateral

Selbststii ndigkeitserkliirun g

Ich erkliire, dass ich die vorliegende Arbeit selbststiindig und unter Verwendungder angegebenen Hilfsmiuel, persdnlichen Mitteilungen und Quellen angefertigthabe.

lena,20MArz202l

//'ortqz ll! L" v

Elham Moayedi

113